Download as pdf or txt
Download as pdf or txt
You are on page 1of 1714

Michael Coey

Stuart S. P. Parkin
Editors

Handbook of
Magnetism
and Magnetic
Materials
Handbook of Magnetism and Magnetic
Materials
J. M. D. Coey • Stuart S. P. Parkin
Editors

Handbook of Magnetism
and Magnetic Materials
With 618 Figures and 157 Tables

123
Editors
J. M. D. Coey Stuart S. P. Parkin
School of Physics Max Planck Institute of Microstructure Physics
Trinity College Halle (Saale)
Dublin Germany
Ireland

ISBN 978-3-030-63208-3 ISBN 978-3-030-63210-6 (eBook)


ISBN 978-3-030-63209-0 (print and electronic bundle)
https://doi.org/10.1007/978-3-030-63210-6

© Springer Nature Switzerland AG 2021


All rights are reserved by the Publisher, whether the whole or part of the material is concerned, specif-
ically the rights of translation, reprinting, reuse of illustrations, recitation, broadcasting, reproduction
on microfilms or in any other physical way, and transmission or information storage and retrieval,
electronic adaptation, computer software, or by similar or dissimilar methodology now known or
hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
The publisher, the authors, and the editors are safe to assume that the advice and information in this book
are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or
the editors give a warranty, expressed or implied, with respect to the material contained herein or for any
errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional
claims in published maps and institutional affiliations.

This Springer imprint is published by the registered company Springer Nature Switzerland AG.
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Preface

Magnetism is a natural phenomenon that arouses curiosity in people of all ages.


Electromagnetism, a mainstay of the industrial revolution, supports urban life and
communications everywhere, served by soft magnetic materials that guide and
concentrate magnetic flux. Permanent magnets are now ubiquitous flux generators,
enabling electric mobility, robotics and energy conversion in a range from μW to
MW. A handful of about a dozen optimized bulk functional magnetic materials
address well over 90% of practical applications.
The ability to pattern magnetic thin films has transformed our subject. Progres-
sively scaled to nanometer dimensions, tiny magnetic regions store binary data,
which forms the basis of today’s digital world. Their stray fields are detected using
minute and exquisitely sensitive magnetic field sensors formed from atomically
engineered multi-layer stacks of magnetic thin films that are the first and, still
today, the most important crop product of spin electronics. Spintronics, especially,
concerns the generation, manipulation and control of the spin angular momentum,
which is the source of the electron’s magnetism. Spin-polarized electrical currents
or main pure spin currents with no net charge flow can be used to excite or switch
the direction of magnetization of magnetic nano-elements. This has opened the door
to a range of magnetic devices with properties that go beyond those of charge-based
electronics. There are new prospects for memory, storage and computation that
are fundamentally spin based. The emerging field of chiral spintronics combines
fundamental aspects of chirality, spin and topology.
On a more fundamental level, although the theoretical foundations of magnetism
in relativity and quantum mechanics were established a century ago, the behaviour
of strongly correlated electrons in solids is an unfailing source of surprises
for physicists and chemists, materials scientists and engineers. Model magnetic
materials can be created to exhibit an astonishing range of physical properties,
and increasingly we are learning how to tailor them to suit a particular practical
application or theoretical model.
The shift of emphasis from bulk, functional magnets to thin films has transformed
the range of elements we can use in our materials. Practically, any stable element in
the periodic table can now be pressed into service, because the quantities needed in
a device are so minute. A billion thin film devices each needing a few nanograms
of some new magnetic material consume just a few grams of an unrecoverable
resource.
v
vi Preface

This handbook aims to offer a broad perspective on the state of the art in
magnetism and magnetic materials. The discovery and dissemination of reliable
knowledge about the natural world is a complex process that depends on interactions
of individuals with shared values and presumptions. Information is the primary
product of their endeavour. It is contained in in papers, patents, reviews, handbooks
monographs and textbooks. This is a perpetual work in progress. Knowledge
percolates through this sequence, taking ever-more digestible and definitive forms
as it is consolidated or eliminated. Now information technology is facilitating this
dynamic. Whereas papers are replaced by more up-to-date papers with new sets
of references to trace their pedigree and textbooks may be updated perhaps after
10 years, handbooks are compendia of information that need updating on a shorter
timescale. This was impractical within the constraints of traditional publication, but
the greater flexibility of electronic publication now opens the possibility for authors
to update their contributions as time passes, and perspectives shift.
The book’s 34 chapters are organized into four parts. After an introduction to the
history and basic concepts in the field, there follow 12 chapters covering the funda-
mentals of solid state magnetism, and the phenomena related to collective magnetic
order. Eight chapters are then devoted to the main classes of magnetic materials –
elements, metallic compounds, oxides and other nonmetallic compounds, thin films,
nanoparticles and artificially engineered materials. Another six chapters treat the
methods for preparing and characterizing magnetic materials, and the final part is
devoted to some major applications.
No fewer than 85 authors have contributed to this handbook. It has taken
longer than we originally anticipated, and the patience of the early responders is
sincerely appreciated. The format for subsequent updating of the electronic text is
by individual chapter, which will avoid such difficulty in the future.
We are grateful to the staff at Springer, Claus Ascheron for initiating the
project, Werner Skolaut for his patience and encouragement, and Barbara Wolf for
efficiently bringing the handbook to hand.

October 2021 J. M. D. Coey


Stuart S. P. Parkin
Contents

Volume 1

Part I Fundamentals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1 History of Magnetism and Basic Concepts . . . . . . . . . . . . . . . . . . . . . 3


J. M. D. Coey
2 Magnetic Exchange Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
Ralph Skomski
3 Anisotropy and Crystal Field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
Ralph Skomski, Priyanka Manchanda, and Arti Kashyap
4 Electronic Structure: Metals and Insulators . . . . . . . . . . . . . . . . . . . . 187
Hubert Ebert, Sergiy Mankovsky, and Sebastian Wimmer
5 Quantum Magnetism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 261
Gabriel Aeppli and Philip Stamp
6 Spin Waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281
Sergej O. Demokritov and Andrei N. Slavin
7 Micromagnetism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 347
Lukas Exl, Dieter Suess, and Thomas Schrefl
8 Magnetic Domains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 391
Rudolf Schäfer
9 Magnetotransport . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 435
Michael Ziese
10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin
Films . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 477
Mark L. M. Lalieu and Bert Koopmans
11 Magnetostriction and Magnetoelasticity . . . . . . . . . . . . . . . . . . . . . . . 549
Dirk Sander

vii
viii Contents

12 Magnetoelectrics and Multiferroics . . . . . . . . . . . . . . . . . . . . . . . . . . . 595


Jia-Mian Hu and Long-Qing Chen
13 Magnetism and Superconductivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . 625
Ilya M. Eremin, Johannes Knolle, and Roderich Moessner

Part II Magnetic Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 657

14 Magnetism of the Elements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 659


Plamen Stamenov
15 Metallic Magnetic Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 693
J. Ping Liu, Matthew Willard, Wei Tang, Ekkes Brück, Frank de
Boer, Enke Liu, Jian Liu, Claudia Felser, Gerhard Fecher, Lukas
Wollmann, Olivier Isnard, Emil Burzo, Sam Liu, J. F. Herbst,
Fengxia Hu, Yao Liu, Jirong Sun, Baogen Shen, and Anne de
Visser
16 Metallic Magnetic Thin Films . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 809
D. Wu and X.-F. Jin

Volume 2
17 Magnetic Oxides and Other Compounds . . . . . . . . . . . . . . . . . . . . . . . 847
J. M. D. Coey
18 Dilute Magnetic Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 923
Alberta Bonanni, Tomasz Dietl, and Hideo Ohno
19 Single-Molecule Magnets and Molecular Quantum Spintronics . . . 979
Gheorghe Taran, Edgar Bonet, and Wolfgang Wernsdorfer
20 Magnetic Nanoparticles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1011
Sara A. Majetich
21 Artificially Engineered Magnetic Materials . . . . . . . . . . . . . . . . . . . . 1047
Christopher H. Marrows

Part III Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1081

22 Magnetic Fields and Measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . 1083


Oliver Portugall, Steffen Krämer, and Yurii Skourski
23 Material Preparation and Thin Film Growth . . . . . . . . . . . . . . . . . . . 1153
Amilcar Bedoya-Pinto, Kai Chang, Mahesh G. Samant, and
Stuart S. P. Parkin
24 Magnetic Imaging and Microscopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1203
Robert M. Reeve, Hans-Joachim Elmers, Felix Büttner, and
Mathias Kläui
Contents ix

25 Magnetic Scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1255


Jeffrey W. Lynn and Bernhard Keimer
26 Electron Paramagnetic and Ferromagnetic Resonance . . . . . . . . . . . 1297
David Menard and Robert Barklie
27 Magnetization Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1333
Andrew D. Kent, Hendrik Ohldag, Hermann A. Dürr, and
Jonathan Z. Sun

Part IV Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1367

28 Permanent Magnet Materials and Applications . . . . . . . . . . . . . . . . . 1369


Karl-Hartmut Müller, Simon Sawatzki, Roland Gauß and
Oliver Gutfleisch
29 Soft Magnetic Materials and Applications . . . . . . . . . . . . . . . . . . . . . . 1435
Frédéric Mazaleyrat
30 Magnetocaloric Materials and Applications . . . . . . . . . . . . . . . . . . . . 1489
Karl G. Sandeman and So Takei
31 Magnetic Sensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1527
Myriam Pannetier-Lecoeur and Claude Fermon
32 Magnetic Memory and Logic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1553
Wei Han
33 Magnetochemistry and Magnetic Separation . . . . . . . . . . . . . . . . . . . 1593
Peter Dunne
34 Magnetism and Biology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1633
Nora M. Dempsey

Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1679
About the Editors

Michael Coey was born in Belfast in 1945. He stud-


ied physics at Cambridge, and then taught English
and physics at the Sainik School, Balachadi (Gujarat).
There he read Allan Morrish’s Physical Principles of
Magnetism from cover to cover (while recovering from
jaundice) before moving to Canada in 1968 to join Mor-
rish’s group at the University of Manitoba for a PhD on
Mõssbauer spectroscopy of iron oxides. He has worked
on magnetism ever since – a life of paid play. After
graduating in 1971, he joined Benoy Chakraverty’s
group at the CNRS in Grenoble as a postdoc with a
letter of appointment signed by Louis Néel. Entering
the CNRS the following year, he worked on the metal-
insulator as well as the magnetism of amorphous solids
and natural minerals. In France, he built the network
of collaborators which sustained much of his career.
On a sabbatical with Stefan von Molnar at the IBM
Research Center at Yorktown Heights, he learned about
magneto-transport and the crystal field. Then, in 1979,
he moved to Ireland as a lecturer at Trinity College
Dublin and set about establishing a magnetism research
group in a venerable but woefully underfunded Physics
Department. Luckily, support from the EU substitu-
tion programme enabled him to begin research on
melt-spun magnetic glasses. Following the discovery
of Nd2 Fe14 B permanent magnets in 1984, he and
colleagues from Grenoble, Birmingham and Berlin
launched the Concerted European Action on Magnets.

xi
xii About the Editors

CEAM blossomed into an informal association of 90


academic and industrial research institutes interested in
every aspect of the properties, processing and appli-
cations of rare-earth iron permanent magnets. He and
his student Sun Hong discovered the interstitial nitride
magnet Sm2 Fe17 N3 in 1990. The group investigated
other rare-earth intermetallic compounds, as well as
magnetic oxide films produced by pulsed-laser deposi-
tion. During this period, he and David Hurley started up
Magnetic Solutions to develop innovative applications
of permanent magnets.
The scientific landscape in Ireland was transformed
by the establishment of Science Foundation Ireland in
2000, given the mission of developing competitive sci-
entific research in Ireland with a budget to match. His
group were able to develop a programme in thin film
magnetism and spin electronics, producing Europe’s
first magnetic tunnel junctions to exhibit 200 % tun-
nel magnetoresistance. Later they discovered the first
zero-moment ferrimagnetic half-metal and explored the
garden of magneto-electrochemistry. Michael coey was
a promotor of CRANN, Ireland’s nanoscience research
centre, and the Science Gallery, now an international
franchise, was his brainchild. Together with Dominique
Givord, he launched the Joint European Magnetic Sym-
posia (JEMS) and, while chair of C9, the IUPAP Mag-
netism Committee, inaugurated the Néel medal that is
awarded triennially at the International Conference on
Magnetism. The 2015 JEMS meeting in Dublin saw a
reunion of many of his 60 PhD students, from all over
the world. Together they have published many papers.
Books include Magnetic Glasses, 1984 (with Kishin
Moorjani): Permanent Magnetism, 1999 (with Ralph
Skomski): and Magnetism and Magnetic Materials,
2010. Honours include Fellowship of the Royal Soci-
ety, International membership of the National Academy
of Sciences, a Fulbright fellowship, a Humboldt Prize,
the Gold Medal of the Royal Irish Academy and the
2019 Born Medal. He has enjoyed visiting profes-
sorships at the University of Strasbourg, the National
University of Singapore and Beihang University in
Beijing.
Michael Coey married Wong May, a writer, in 1973;
they have two sons and a grand-daughter.
About the Editors xiii

Stuart S. P. Parkin is a director of the Max Planck


Institute of Microstructure Physics, Halle, Germany,
and an Alexander von Humboldt Professor, Mar-
tin Luther University, Halle-Wittenberg. His research
interests include spintronic materials and devices for
advanced sensor, memory and logic applications, oxide
thin-film heterostructures, topological metals, exotic
superconductors, and cognitive devices. Stuart’s dis-
coveries in spintronics enabled a more than 10,000-
fold increase in the storage capacity of magnetic disk
drives. For his work that, thereby, enabled the ‘big
data’ world of today. In 2014, he was awarded the
Millennium Technology Award from the Technology
Academy Finland and, most recently, the King Faisal
Prize for Science 2021 for his research into three
distinct classes of spintronic memories. Stuart is a
fellow or member of: The Royal Society, the Royal
Academy of Engineering, the National Academy of
Sciences, the National Academy of Engineering, the
German National Academy of Science – Leopoldina,
The Royal Society of Edinburgh, The Indian Academy
of Sciences, and TWAS – The academy of sciences
for the developing world. Stuart is also a fellow of
the American Physical Society: the Institute of Elec-
trical and Electronics Engineers (IEEE) the Institute
of Physics, London: the American Association for the
Advancement of Science (AAAS); and the Materials
Research Society. Stuart has published more than 600
papers and has more than 121 issued patents. His h
factor is 120. Clarivate Analytics named him a Highly
Cited Researcher in 2018, 2019, 2020 and 2021.
Stuart’s numerous awards include the American
Physical Society International Prize for New Mate-
rials (1994); the Europhysics Prize for Outstanding
Achievement in Solid State Physics (1997); the 2009
IUPAP Magnetism Prize and Néel Medal; the 2012
von Hippel Award – Materials Research Society; the
2013 Swan Medal – Institute of Physics; an Alexander
von Humboldt Professorship – International Award for
Research (2014); and ERC Advanced Grant – SOR-
BET (2015). Stuart has been a distinguished visiting
professor at several universities worldwide including:
National University of Singapore; National Taiwan
University; National Yunlin University of Science and
Technology, Taiwan; Eindhoven University of Tech-
xiv About the Editors

nology, The Netherlands; KAIST, Korea; and Univer-


sity College London. Stuart has been awarded four
honorary doctorates by: RWTH Aachen University
(2007), Eindhoven University of Technology (2008),
The University of Regensburg (2011), and Technische
Universität Kaiserslautern, Germany (2013).
Prior to being appointed to the Max Planck Society,
Stuart had spent a large part of his career with IBM
Research at the San Jose Research Laboratory, which
became the Almaden Research Center when it moved
to a new campus. Stuart was appointed an IBM Fellow,
IBM’s highest technical honour, by IBM’s chairman,
Louis Gerstner in 1999. He received his BA physics
and theoretical physics (1977), an MA, and his PhD
(1980) from the University of Cambridge. He was a stu-
dent at Trinity College, Cambridge, where he received
an entrance scholarship (1974), a senior scholarship
(1975), a research scholarship (1977) and was elected
a research fellow (1979). In 2014, he became an hon-
orary fellow. Stuart received a Royal Society European
Exchange Fellowship to carry out postdoctoral research
at the Laboratoire de Physique des Solides, Université
Paris-Sud, France, in 1980–1981 and an IBM World
Trade Fellowship to carry out research at IBM in San
Jose.
Contributors

Gabriel Aeppli Physics Department (ETHZ), Institut de Physique (EPFL) and


Photon Science Division (PSI), ETHZ, EPFL and PSI, Zürich, Lausanne and
Villigen, Switzerland
Robert Barklie School of Physics, Trinity College, Dublin, Ireland
Amilcar Bedoya-Pinto Max Planck Institute of Microstructure Physics, Halle
(Saale), Germany
Alberta Bonanni Institut für Halbleiter- und Festkörperphysik, Johannes Kepler
University, Linz, Austria
Edgar Bonet Néel Institute, CNRS, Grenoble, France
Ekkes Brück Delft University of Technology, Delft, The Netherlands
Emil Burzo Babes-Bolyai University, Romania, Cluj-Napoca, Romania
Felix Büttner Helmholtz-Zentrum Berlin für Materialien und Energie, Berlin,
Germany
Kai Chang Beijing Academy of Quantum Information Sciences, Beijing, China
Long-Qing Chen Materials Research Institute, and Department of Materials Sci-
ence and Engineering, The Pennsylvania State University, University Park, PA, USA
Michael Coey School of Physics, Trinity College, Dublin, Ireland
Frank de Boer University of Amsterdam, Amsterdam, The Netherlands
Anne de Visser Van der Waals-Zeeman Institute, University of Amsterdam,
Amsterdam, The Netherlands
Sergej O. Demokritov Institute for Applied Physics and Center for Nanotechnol-
ogy, University of Muenster, Muenster, Germany
Nora M. Dempsey Institut Néel, CNRS & Université Grenoble Alpes, Grenoble,
France

xv
xvi Contributors

Tomasz Dietl International Research Centre MagTop, Institute of Physics, Polish


Academy of Sciences, Warsaw, Poland
WPI Advanced Institute for Materials Research, Tohoku University, Sendai, Japan
Peter Dunne Institut de Physique et de Chimie des Matériaux de Stasbourg,
Strasbourg, France
Hermann A. Dürr Department of Physics and Astronomy, Uppsala University,
Uppsala, Sweden
Hubert Ebert München, Department Chemie, Ludwig-Maximilians-Universität,
München, Germany
Hans-Joachim Elmers Institute of Physics, Johannes Gutenberg University Mainz,
Mainz, Germany
Ilya M. Eremin Institut für Theoretische Physik III, Ruhr-Universität Bochum,
Bochum, Germany
Lukas Exl University of Vienna Research Platform MMM Mathematics – Mag-
netism – Materials, University of Vienna, and Wolfgang Pauli Institute, Wien,
Austria
Gerhard Fecher Max-Planck-Institute für Chemische Physik fester Stoffe, Dres-
den, Germany
Claudia Felser Max-Planck-Institute für Chemische Physik fester Stoffe, Dresden,
Germany
Claude Fermon Service de Physique de l’Etat Condensé, DRF/IRAMIS/SPEC
CNRS UMR 3680 CEA Saclay, Gif sur Yvette, France
Roland Gauß EIT RawMaterials GmbH, Berlin, Germany
Oliver Gutfleisch Technische Universität Darmstadt, Materialwissenschaft, Darm-
stadt, Germany
Wei Han International Center for Quantum Materials, School of Physics, Peking
University, Beijing, China
J. F. Herbst Research & Development, General Motors R&D Center, Warren,
MI, USA
Fengxia Hu Institute of Physics, Chinese Academy of Sciences, Beijing, China
Jia-Mian Hu Department of Materials Science and Engineering, University of
Wisconsin-Madison, Madison, WI, USA
Olivier Isnard Institute Néel and Université Grenoble Alpes, Grenoble, France
X.-F. Jin Department of Physics and State Key Laboratory of Surface Physics,
Fudan University, Shanghai, People’s Republic of China
Arti Kashyap IIT Mandi, Mandi, HP, India
Contributors xvii

Bernhard Keimer Max-Planck Institute for Solid State Research, Stuttgart,


Germany
Andrew D. Kent Center for Quantum Phenomena, Department of Physics, New
York University, New York, NY, USA
Mathias Kläui Institute of Physics, Johannes Gutenberg University Mainz, Mainz,
Germany
Johannes Knolle Blackett Laboratory, Imperial College London, London, UK
Bert Koopmans Department of Applied Physics, Eindhoven University of Tech-
nology, Eindhoven, The Netherlands
Steffen Krämer LNCMI-CNRS (UPR3228), EMFL, Univ. Grenoble Alpes, INSA
Toulouse, Univ. Toulouse 3, Grenoble, France
Mark L. M. Lalieu Department of Applied Physics, Eindhoven University of
Technology, Eindhoven, The Netherlands
Enke Liu Institute of Physics, Chinese Academy of Sciences, Beijing, China
J. Ping Liu University of Texas at Arlington, Arlington, TX, USA
Jian Liu Ningbo Institute of Materials Technology and Engineering, Chinese
Academy of Sciences, Ningbo, China
Sam Liu University of Dayton, Dayton, OH, USA
Yao Liu Institute of Physics, Chinese Academy of Sciences, Beijing, China
Jeffrey W. Lynn NIST Center for Neutron Research, National Institute of Stan-
dards and Technology, Gaithersburg, MD, USA
Sara A. Majetich Physics Department, Carnegie Mellon University, Pittsburgh,
PA, USA
Priyanka Manchanda Howard University, Washington, DC, USA
Sergiy Mankovsky München, Department Chemie, Ludwig-Maximilians-
Universität, München, Germany
Christopher H. Marrows School of Physics and Astronomy, University of Leeds,
Leeds, United Kingdom
Frédéric Mazaleyrat SATIE, CNRS, École Normale Supérieure Paris-Saclay,
Gif-sur-Yvette, France
David Menard Department of Engineering Physics, Polytechnique Montreal, Mon-
tréal, QC, Canada
Roderich Moessner Max-Planck Institut für Physik komplexer Systeme, Dresden,
Germany
Karl-Hartmut Müller IFW Dresden, Institute for Metallic Materials, Dresden,
Germany
xviii Contributors

Hendrik Ohldag Advanced Light Source, Lawrence Berkeley National Laboratory,


Berkeley, CA, USA
Department of Physics, University of California Santa Cruz, Santa Cruz, CA,
USA
Department of Materials Science, Stanford University, Stanford, CA, USA
Hideo Ohno WPI Advanced Institute for Materials Research, Tohoku University,
Sendai, Japan
Laboratory for Nanoelectronics and Spintronics, Research Institute of Electrical
Communication, Tohoku University, Sendai, Japan
Center for Spintronics Integrated System, Tohoku University, Sendai, Japan
Center for Innovative Integrated Electronic Systems, Tohoku University, Sendai,
Japan
Center for Science and Innovation in Spintronics (Core Research Cluster), Tohoku
University, Sendai, Japan
Center for Spintronics Research Network, Tohoku University, Sendai, Japan
Myriam Pannetier-Lecoeur Service de Physique de l’Etat Condensé, DRF/
IRAMIS/SPEC CNRS UMR 3680 CEA Saclay, Gif sur Yvette, France
Stuart S. P. Parkin Max Planck Institute of Microstructure Physics, Halle (Saale),
Germany
Oliver Portugall LNCMI-CNRS (UPR3228), EMFL, Univ. Grenoble Alpes, INSA
Toulouse, Univ. Toulouse 3, Toulouse, France
Robert M. Reeve Institute of Physics, Johannes Gutenberg University Mainz,
Mainz, Germany
Mahesh G. Samant IBM Research, San Jose, CA, USA
Karl G. Sandeman Department of Physics, Brooklyn College of the City
University of New York, Brooklyn, NY, USA
The Physics Program, The Graduate Center, CUNY, New York, NY, USA
Dirk Sander Max Planck Institute of Microstructure Physics, Halle, Germany
Simon Sawatzki Technische Universität Darmstadt, Materialwissenschaft,
Darmstadt, Germany
Vacuumschmelze GmbH & Co.KG, Hanau, Germany
Rudolf Schäfer Institute for Metallic Materials, Leibniz Institute for Solid State
and Materials Research (IFW) Dresden, Dresden, Germany
Institute for Materials Science, Dresden University of Technology, Dresden,
Germany
Contributors xix

Thomas Schrefl Christian Doppler Laboratory for Magnet Design Through Physics
Informed Machine Learning, Department of Integrated Sensor Systems, Danube
University Krems, Wiener Neustadt, Austria
Baogen Shen Institute of Physics, Chinese Academy of Sciences, Beijing, China
Ralph Skomski University of Nebraska, Lincoln, NE, USA
Yurii Skourski Hochfeld-Magnetlabor Dresden (EMFL-HLD), Helmholtz-
Zentrum Dresden-Rossendorf, Dresden, Germany
Andrei N. Slavin Department of Physics, Oakland University, Rochester, MI, USA
Plamen Stamenov School of Physics and CRANN, Trinity College, University of
Dublin, Dublin, Ireland
Philip Stamp Pacific Institute of Theoretical Physics, University of British
Columbia, Vancouver, BC, Canada
Dieter Suess University of Vienna Research Platform MMM Mathematics – Mag-
netism – Materials, and Physics of Functional Materials, Faculty of Physics,
University of Vienna,Wien, Austria
Jirong Sun Institute of Physics, Chinese Academy of Sciences, Beijing, China
Jonathan Z. Sun IBM T. J. Watson Research Center, Yorktown Heights, NY, USA
So Takei The Physics Program, The Graduate Center, CUNY, New York, NY, USA
Department of Physics, Queens College of the City University of New York,
Flushing, NY, USA
Wei Tang Materials Science and Engineering, Ames Laboratory, Ames, IA, USA
Gheorghe Taran Physikalisches Institute, KIT, Karlsruhe, Germany
Wolfgang Wernsdorfer Physikalisches Institute, KIT, Karlsruhe, Germany
Matthew Willard Materials Science and Engineering, Case Western Reserve
University, Cleveland, OH, USA
Sebastian Wimmer München, Department Chemie, Ludwig-Maximilians-
Universität, München, Germany
Lukas Wollmann Max-Planck-Institute für Chemische Physik fester Stoffe,
Dresden, Germany
D. Wu National Laboratory of Solid State Microstructures and Department of
Physics, Nanjing University, Nanjing, People’s Republic of China
Michael Ziese Fakultät für Physik und Geowissenschaften, Universität Leipzig,
Leipzig, Germany
Part I
Fundamentals
History of Magnetism and Basic Concepts
1
J. M. D. Coey

Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
Early History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
The Compass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
The Emergence of Modern Science . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
The Electromagnetic Revolution [9] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
Magnetostatics and Classical Electrodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
The Earth’s Magnetic Field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
The Properties of Ferromagnets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
Magnetism of the Electron . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
The Demise of Classical Physics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
Magnetic Phenomenology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
Micromagnetism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
Magnetic Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
Magnetic Oxides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
Intermetallic Compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
Model Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
Amorphous Magnets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
Magnetic Fine Particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
Magnetic Recording . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
Methods of Investigation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
Materials Preparation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
Experimental Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
Computational Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
Spin Electronics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
Appendix: Units . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

J. M. D. Coey ()
School of Physics, Trinity College, Dublin, Ireland
e-mail: jcoey@tcd.ie

© Springer Nature Switzerland AG 2021 3


J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_1
4 J. M. D. Coey

Abstract

Magnetism is a microcosm of the history of science over more than two


millennia. The magnet allows us to manipulate a force field which has catalyzed
an understanding of the natural world that launched three revolutions. First
came the harnessing of the directional nature of the magnetic force in the
compass that led to the exploration of the planet in the fifteenth century.
Second was the discovery of the relation between electricity and magnetism that
sparked the electromagnetic revolution of the nineteenth century. Third is the
big data revolution that is currently redefining human experience while radically
transforming social interactions and redistributing knowledge and power.
The emergence of magnetic science demanded imagination and observational
acuity, which led to the theory of classical electrodynamics. The magnetic field is
associated with electric currents and the angular momentum of charged particles
in special materials. Our current understanding of the magnetism of electrons
in solids is rooted in quantum mechanics and relativity. Yet only since about
1980 has fundamental theory underpinned rational design of new functional
magnetic materials and the conception of new spin electronic devices that can
be reproduced on ever smaller scales, leading most notably to the disruptive,
60-year exponential growth of magnetic information storage. The development
of new magnetic concepts, coupled with novel materials, device and machine
designs has become a rich source of technical innovation.

Introduction

The attraction of ferrous objects to a permanent magnet has been a source of wonder
since the Iron Age. Feeble magnets are widespread in nature in the form of rocks
known as lodestones, which are rich in magnetite, an oxide mineral with ideal
formula Fe3 O4 . Rocky outcrops eventually get magnetized by huge electric currents
when lightning strikes, and these natural magnets were known and studied in ancient
Greece, Egypt, China, and Mesoamerica. Investigations of magnetic phenomena led
to the invention of steel magnets – needles and horseshoes – then electromagnets and
eventually the panoply of hard and soft materials that support the modern magnetics
industry. Magnetism in a rare example of a science with recorded history goes back
well over 2000 years [1, 2].
Theory and practice have been loose partners for most of that time. What people
are able to see and rationalize is inevitably conditioned by a priori philosophical
beliefs about the world. The scientific method of critically interrogating nature
by experimentation and then amassing and exchanging data and ideas among the
community of the curious came to be established only gradually. Mathematics
emerged as the supporting scaffold of natural philosophy in Europe in the seven-
teenth century, when precisely formulated natural laws and explanations began to
take root. Nevertheless, most of the progress that has been made in magnetism in
the past – from the discovery of horseshoe magnets or electromagnetic induction
1 History of Magnetism and Basic Concepts 5

to the development of Alnico – was based on intuition and experience, rather than
formal theory. That situation is changing.
The discovery of the electron in the closing years of the nineteenth century
impelled the great paradigm shift from classical to modern physics. Magnetism,
however familiar and practically important it had become, was fundamentally
incomprehensible in classical terms. Charged particles were theoretically expected
to exhibit no magnetism of any kind. It took 25 years and the insights of quantum
mechanics and relativity to resolve that conundrum. Magnetism then went on to play
a key role in clarifying basic concepts in condensed matter physics and Earth science
over the course of the twentieth century. Now it is a key player in the transformative
information technology of the twenty-first century.

Early History

Aristotle attributed the first reflections on the nature of magnetic attraction to Thales,
the early Greek philosopher and mathematician who was born in Miletus in Asia
Minor in 624 BC. Thales was an animist who credited the magnet with a soul,
on account of its ability to create movement, by attraction. This curious idea was
to linger until the seventeenth century. The magnet itself is believed to be named
after Magnesia, a city in Lydia in Asia Minor that was a good source of lodestone.
In the fifth century BC, when Empedokles postulated the existence of the four
elements – Earth, water, air, and fire – magnetism was associated with air. Special
effluvia somehow passing through the invisible pores in magnetic material were
invoked to explain the phenomenon, a theory echoed much later by Descartes in
a mechanistic picture that finally laid the magnet’s soul to rest. The Roman poet
Lucretius writing in the first century BC mentions magnetic induction (the ability
of a magnet to induce magnetism in pieces of nonmagnetic iron) and for the first
time notes the ability of magnets not just to attract but also to repel one another.
The Greek approach of developing a philosophical framework into which natural
observations were expected to fit was not conducive to open-minded exploration of
the natural world.

The Compass

The Chinese approach to the magnet was more practical. Their magnetism was
initially linked to practical concerns of geomancy and divination [3]. The art of
adapting the residences of the living and the tombs of the dead to harmonize
with local currents of the cosmic breath demanded knowledge of its direction. A
south-pointer consisting of a carved lodestone spoon that was free to rotate on a
polished baseplate (Fig. 1) was already in use at the time of Lucretius and may have
originated hundreds of years earlier. An important discovery, attributed to Zeng
Gongliang in 1064, was that iron could acquire a thermoremanent magnetization
when rapidly cooled from red heat in the Earth’s magnetic field. A short step
6 J. M. D. Coey

Fig. 1 Magnetic direction finders. (a) Baseplate and lodestone spoon of the south-pointer used in
China from about the first century BC (Needham, courtesy of Cambridge University Press). (b) A
Chinese floating compass from 1044. (c) Fifteenth-century Chinese and (d) Portuguese mariners’
compasses. (Boorstin, courtesy of Editions Robert Laffont)

led to the suspended compass needle, which was described by Shen Kuo around
1088, together with declination, the deviation of the needle from a north-south axis.
Floating compasses had also been developed by this time, often in the form of an
iron fish made to float in a bowl of water.
The compass appeared about a century later in Europe, where it was first
described by Alexander Neckam in 1190. The direction-finding ability of the
magnetic needle or fish was also exploited by Arabs and Persians from the thirteenth
century, both for navigation and to determine the sacred direction of Mecca [4].
Compasses (Fig. 1) were the enabling technology for the great voyages of discovery
of the fifteenth century, bringing the Ming admiral Cheng Ho to the coasts of Africa
in 1433 and Christopher Columbus (who rediscovered declination) to America in
1492, where he landed on the continent where the Olmecs may once have displayed
a knowledge of magnetism in their massive stone carvings of human figures and sea
turtles dating from the second millennium BC.
1 History of Magnetism and Basic Concepts 7

Before long, the landmasses and oceans of our planet were mapped and explored.
According to Francis Bacon, writing in Novum Organum in 1620 [5], the magnetic
compass was one of three things, along with printing and gunpowder had “changed
the whole face and state of things throughout the world.” All three were originally
Chinese inventions. The compass helped to provide us with an image of the planet
we inhabit. This was the first of three occasions when magnetism changed the world.

The Emergence of Modern Science

A landmark in the history of magnetism in Europe was the work of the French
crusader monk Petrus Peregrinus. His tract Epistola de Magnete [6] recounts
experiments with floating pieces of lodestone and carved lodestone spheres called
terella, which he wrote up in Southern Italy during the 1269 siege of Lucera. He
describes how to find the poles of a magnet and relates magnetic attraction to
the celestial sphere. The same origin had long been associated with the magnet’s
directional property in China [3]; we should not forget that before electric light,
people were acutely aware of the stars and scrutinized them keenly. Peregrinus’s
tract included an ingenious proposal for a magnetic perpetual motion device – a
theme that has been embraced by charlatans throughout the ages, right up to the
present day.
Much credit for the inauguration of the experimental method in a recognizably
modern form belongs to William Gilbert. Physician to the English Queen Elizabeth
I, Gilbert personally conducted a series of experiments on terellas, which led him to
proclaim that the Earth itself was a great magnet. The lodestone or steel magnets
aligned themselves not with the celestial sphere, but with the Earth’s poles. He
induced magnetism by cooling iron in the Earth’s field and then destroyed it by
heating or hammering. Gilbert was at pains to debunk the millennial accretion of
superstition that clung to the magnet, confidently advocating in a robust polemical
style reliance on the evidence of one’s own eyes. He described his investigations in
his masterwork De Magnete, published in 1600 [7]. It is arguably the first modern
scientific text.
Subsequent developments were associated with improvements in navigation and
the prestige of the great voyages of discovery. Gilbert’s theories dominated the
seventeenth century up until Edmond Halley’s 1692 shell model for the Earth’s
magnetic structure, which strongly influenced compass technology and navigation.
Naval interests were the principal drivers of magnetic research during this period,
and Halley was sponsored by the British Navy to survey and prepare charts of
the Earth’s magnetic field in the North and South Atlantic oceans (Fig. 2), This
was in the vain hope of addressing the pressing longitude problem, by pinpointing
magnetically the position of a vessel on the Earth’s surface.
The following century was marked by the professionalization of natural philos-
ophy (as physical science was then known in Europe) [8]. Accordingly, the natural
philosopher with his mantle of theory was rewarded with social status, access to
public funding, and credibility beyond that extended to artisans on the one hand and
8 J. M. D. Coey

Fig. 2 A section of Halley’s world chart of magnetic variation published in 1700

quacks on the other, such as the colorful Anton Mesmer, who propagated theories of
animal magnetism in his salon in Paris or James Graham with his royal Patagonian
magnetic bed for nightly rental in a fashionable London townhouse. The English
entrepreneur Gowin Knight, representative of a new breed of natural philosopher,
greatly improved the quality of bar magnets and compasses, coupling scientific
endeavor with manufacturing enterprise and a keen sense of intellectual property.
An outstanding technical breakthrough of the eighteenth century was the 1755
discovery by the Swiss blacksmith Johann Dietrich that the horseshoe was an ideal
1 History of Magnetism and Basic Concepts 9

shape for a steel magnet [1]. His invention, a clever practical solution to the age-old
problem of self-demagnetization in bar magnets, was enthusiastically promoted by
his mentor, the Swiss applied mathematician Daniel Bernoulli, who garnered most
of the credit.

The Electromagnetic Revolution [9]

The late eighteenth century in Europe was a time of great public appetite for lectures
and demonstration of the latest scientific discoveries, not least in electricity and
magnetism. This effervescent age witnessed rapid developments in the harnessing
of electricity, with the 1745 invention of the Leyden jar culminating in Alessandro
Volta’s 1800 invention of the voltaic cell. Analogies between electrostatics and
magnetism were tantalizing, but the link between them proved elusive.

Magnetostatics and Classical Electrodynamics

The torsion balance allowed Charles-Augustin de Coulomb to establish in 1785


the quantitative inverse square laws of attraction and repulsion between electric
charges, as well as similar laws between analogous magnetic charge or poles that
were supposed to be located near the ends of long magnetized steel needles [2]. The
current convention is that the north and south magnetic poles are negatively and
positively charged, respectively. His image was of pairs of positive and negative
electric and magnetic fluids permeating matter, which became charged if one
of them dominated or polarized if they were spatially separated. Unlike their
electric counterparts, the magnetic fluids were not free to flow and could never
be unbalanced in any piece of magnetic material. Coulomb found that the force F
between two magnetic poles separated by a distance r fell away as 1/r2 . Siméon
Denis Poisson then interpreted Coulomb’s results in terms of a scalar potential
ϕm , analogous to the one he used for static electricity, such that the magnetic field
could be written as H(r) = −∇ϕm . In modern terms, ϕm is measured in amperes,
and H in Am−1 . Magnetic charge qm is measured in Am, and the corresponding
potential ϕm = qm /4πr. The magnetic field due to a charge is H(r) = qm r/4πr3 ,
and Coulomb’s inverse square law for the force between two charges separated by
r is F = μ0 qm qm ’r/4πr3 . Here μ0 is the magnetic constant, 4π 10−7 NA−2 , which
appears whenever the magnetic field H interacts with matter. (Other equivalent ways
of writing the units of μ0 are Hm−1 or TmA−1 .)
In Poisson’s opinion, the practice and teaching of mathematics were the purpose
of life. He developed his mathematical theory of magnetostatics from 1824, which
included the equation that bears his name ∇ 2 ϕm = −ρm , where ρm is the density of
magnetic poles. However, the association of H with a scalar potential is only valid
in a steady state and when no electric currents are present. The coulombian picture
of the origin of magnetic fields was dominant in textbooks until about 1960, and it
persists in popular imagery.
10 J. M. D. Coey

A revolutionary breakthrough in the history of magnetism came on 21st April


1820, with the discovery of the long-sought link between electricity and magnetism.
During a public lecture, the Danish scientist Hans Christian Oersted noticed that a
compass needle was deflected as he switched on an electric current in a copper
wire. His report, published in Latin a few months later, triggered an experimental
frenzy. As soon as the news reached Paris, François Arago (who briefly served as
President of France in 1848) immediately performed an experiment that established
that a current-carrying conducting coil behaved like a magnet. A week after Arago’s
report, André-Marie Ampère presented a paper to the French Academy suggesting
that ferromagnetism in a magnetized body was caused by internal currents flowing
perpendicular to the axis of magnetization and that it should therefore be possible
to magnetize steel needles in a solenoid. Together with Arago, he successfully
demonstrated his ideas in November 1820, showing that current loops and coils
were functionally equivalent to magnets, and he subsequently established the
law of attraction or repulsion between current-carrying wires. Ten days later,
the British scientist Humphrey Davy had similar results. The electromagnet was
invented by William Sturgeon in 1825; within 5 years Joseph Henry had used a
powerful electromagnet in the USA for the first electric telegraph. As early as 1822,
Davy’s assistant Michael Faraday produced the first rudimentary electric motor, and
Ampère envisaged the possibility that the currents causing magnetism in solids were
“molecular” rather than macroscopic in nature.
In formal terms, Ampère’s equivalence between a magnet and a current loop of
area A carrying a current I is expressed as

m = IA (1)

where A is in square meters, I is in amperes, and the magnetic moment m is therefore


in Am2 . Magnetization, defined in a mesoscopic volume V as M = m/V, has units
Am−1 . The direction of m is conventionally related to that of the electric current
by the right-hand rule. At the same time as the experimental work of Ampère
and Arago, Jean-Baptiste Biot and Félix Savart formulated the law expressing the
relation between a current and the field it produces. A current element Iδl generates
a field δH = Iδl × r/4πr3 at a distance r. Integrating around a current loop yields an
expression for the H-field due to the moment m:

H = [3 (m.r) r − m] /4r 3 (2)

The form of the field represented by Eq. (2) and illustrated in Fig. 3 is identical
to that of an electric dipole, so m is often referred to as a magnetic dipole although
we have no evidence for the existence of independent magnetic poles. The dipole
moment is best represented by an arrow in the direction of m, although it is still
commonplace to see the north-seeking and south-seeking poles of a magnet denoted
by the letters N and S. Old habits die hard.
Magnetic moments tend to align with magnetic fields in which they are placed.
The torque on the dipole m is Γ = μ0 m × H, and the corresponding energy of the
1 History of Magnetism and Basic Concepts 11

Fig. 3 Contours of equal magnetic field produced by a magnetic dipole moment m, represented
by the grey arrow

dipole is E = − μ0 m . H. These equations are better written in terms of the more


fundamental magnetic field B, as discussed below; in free space the two are simply
proportional, B = μ0 H, so the torque is

 =m×B (3)

and the corresponding energy is

E = −m.B (4)

The two rival descriptions of magnetization in solids following from the work of
Coulomb or Ampère, based either on magnetic poles or on electric currents, have
colored thinking about magnetism ever since (Fig. 4). The poles have no precise,
independent physical reality; they are fictitious entities that are a mathematically-
convenient way to represent the H-field, which is of critical importance in mag-
netism because it is the local H-field that determines the state of magnetization
of a solid. Currents are closer to reality; electric current loops exist, and they do
act like magnets. Although it is difficult to attribute the intrinsic spin moment of
the electron to a current, the amperian picture of the origin of magnetic fields is
generally adopted in modern textbooks.
Nineteenth-century electromagnetism owed much to the genius of Michael
Faraday. Guided entirely by observation and experiment, with no dependence on
formal theory, he was able to perfect the concept of magnetic field, which he
12 J. M. D. Coey

Fig. 4 Alternative
+++++ σ m+
coulombian (left) and
amperian (right) descriptions
of the magnetization of a
uniformly magnetized
cylinder, with a magnetic
dipole moment m in the
direction represented by the
jms
black arrow; σm ± is the
surface magnetic charge
density, jms is the surface
electric current density
-----
σ m-

described by lines of force [10]. Faraday classified substances in three magnetic


categories. Ferromagnets like iron were spontaneously magnetized and strongly
attracted into a magnetic field; paramagnets were weakly magnetized by a field
and feebly drawn into the regions where the field was strongest; diamagnets, on
the contrary, were weakly magnetized opposite to the field and repelled by it.
Working with an electromagnet, he discovered the law that bears his name and
the phenomenon of electromagnetic induction – that a flow of electricity can be
induced by a changing magnetic field – in 1831. His conviction that a magnetic field
should have some effect on light led to his 1845 discovery of the magneto-optic
Faraday effect – that the plane of polarization of light rotates upon passing through
a transparent medium in a direction parallel to the magnetization of the medium.
The epitome of classical electrodynamics was the set of equations formulated
in 1865 by James Clerk Maxwell, the Scottish theoretician, who had “resolved
to read no mathematics on the subject till he had first read through Faraday’s
‘Experimental Researches in Electricity’.” Maxwell’s magnificent equations for-
mally defined the relationship between electricity, magnetism, and light [11]. As
reformulated by Oliver Heaviside, the equations are a succinct statement of classical
electrodynamics. In the opinion of Richard Feynman, Maxwell’s discovery of the
laws of electrodynamics was the most significant event of the nineteenth century.
The equations in free space are formulated in terms of the fundamental magnetic
and electric fields B and E. Using the international system of SI units adopted in
this Handbook, the equations read:

∇.B = 0
ε0 ∇.E = ρ
(5)
(1/μ0 ) ∇ × B = j + ε0 ∂E/∂t
∇ × E = −∂B/∂t

The first and third equations express the idea that there are no sources of the
magnetic B-field other than time-varying electric fields and electric currents of
1 History of Magnetism and Basic Concepts 13

density j, whereas the second and fourth equations show that the electric field
results from electric charge density ρ and time-varying magnetic fields. Maxwell’s
equations are invariant in a moving frame of reference, although the relative
magnitudes of E and B are altered.
The famous wavelike solutions of these equations in the absence of charges and
currents are electromagnetic waves, which propagate in free space with velocity
c = 1/(ε0 μ0 )1/2 . In SI, the definition of the magnetic constant μ0 is linked to the
fine structure constant. To nine significant figures, it is equal to 4π 10−7 NA−2 . ε0
is then related to the definition of the velocity of light. Heinrich Hertz demonstrated
Maxwell’s electromagnetic waves experimentally in 1888, and he showed that their
behavior was essentially the same as that of light. Hertz could think of no practical
application for his work, yet within a few decades, it had become the basis of radio
broadcasting and wireless communication!
The mechanical effects of electric and magnetic fields were summarized by
Hendrik Lorentz in his expression for the force density FL :

F L = ρE + j × B (6)

The equivalent expression for the force on a particle of charge q moving with
velocity v is f = q(E + v × B).
Two further fields H and D are introduced in the formulation of Maxwell’s
equations in a material medium to circumvent the inaccessibility of the current and
charge distributions in the medium. We have no direct way of measuring the atomic
charges associated with the polarization of a ferroelectric material or the atomic
currents associated with the magnetization of a ferromagnetic material, so we define
H and D in terms of fields created by the measurable free charges ρ and free currents
j, with dipolar contributions from the magnetization M or polarization P of any
magnetic or dielectric material that may be present. The equations now read:

∇.B = 0
∇.D = ρ
(7)
∇ × H = j + ∂D/∂t
∇ × E = −∂B/∂t

They are further simplified in a static situation when the time derivatives are
zero. The new fields are trivially related to B and E in free space since B = μ0 H and
D = ε0 E, but in a material medium, the H-field is defined in terms of the B-field and
the magnetization M (the magnetic moment per unit volume) as H = B/μ0 – M or

B = μ0 (H + M) (8)
14 J. M. D. Coey

H M B
+++++

–––––

B = P0(H + M)

Fig. 5 B, H, and M for a uniformly magnetized ferromagnetic bar. Eq. (8) is represented by the
vector triangle. The H-field can be regarded as originating from a distribution of positive and
negative magnetic charge (south and north magnetic poles) on opposite faces

Likewise D = ε0 (E + P), where P is the electric polarization. To specify a


situation in magnetostatics or electrostatics, any two of the three magnetic or electric
fields are needed. (Magnetization M and polarization P are regarded as vector
fields.) The defining relation between B, H, and M for a uniformly magnetized
ferromagnetic bar is illustrated in Fig. 5. Note that the B-field is solenoidal – the field
lines are continuous with no sources or sinks; it is divergenceless and can therefore
be expressed as the curl of a vector potential A – whereas the H-field is conservative;
it is irrotational provided j is zero and can be expressed as the gradient of a scalar
potential. Outside the magnet, the H-field is called the stray field, but within the
magnet where it is oppositely oriented to M, the name changes to demagnetizing
field. Boundary conditions that B⊥ and H|| are continuous across an interface in
a steady state (j = 0) follow from the first and third of Maxwell’s equations 7. B
is the fundamental magnetic field, because no elementary magnetic poles exist in
nature (∇. B = 0), but it is the local value of H (and perhaps the sample history)
that determines the magnetic state of a solid, including its micromagnetic domain
structure. The H-field acting in a solid is the sum of the applied field H and the
local demagnetizing field Hd created by the solid body itself.
When describing the stray field outside a distribution of magnetization M(r)
in a solid, the coulombian and amperian descriptions are formally equivalent.
The coulombian expression for the magnetic field is obtained by integrating the
expression for the field due to a distribution of a magnetic charge qm per unit volume
ρm = −∇. M in the bulk, and per unit area σm = M. en at the surface, where en is
the unit vector normal to the surface:
1 History of Magnetism and Basic Concepts 15

         
1 ∇ · M r − r 3  M · en r − r  2 
H (r) = − d r + d r (9)
4π V |r − r  |3 S |r − r  |3

This formula gives H(r) both inside and outside the magnetic material. Outside
B(r) = μ0 H(r).
The amperian expression for the magnetic field produced by a distribution of
currents is based on the Biot-Savart expression for the field due to a current element,
including contributions from the current density jm = ∇ × M in the bulk, and
jms = M × en at the surface:
         
μ0 ∇ × M × r − r 3  (M × en ) × r − r  2 
B (r) = d r + d r
4π V |r − r  |3 S |r − r  |3
(10)

This formula gives B(r) both inside and outside the magnetic material. The same
result can be obtained by appropriate integration of Eq. 2 over a magnetization
distribution M(r) [12].
For uniformly magnetized ellipsoids, the demagnetizing field Hd is related to the
magnetization by

H d = −N M (11)

where N is a tensor with unit trace [13]. It reduces to a simple scalar demagnetizing
factor 0 < N < 1 when the magnetization lies along a principal axis of the ellipsoid.
N ≈ 0 for a long needle magnetized along its axis, and N = 1 for a flat plate
magnetized perpendicular to the plane. A sphere has N = 1/3. For any shape
less symmetric than an ellipsoid, the demagnetizing field is nonuniform. There are
useful approximate formulae for square bars and cylinders [14], such as 1/(2n + 1)

and 1/[(4n/ π) + 1], respectively, but they should not obscure the fact that the
demagnetizing field in these shapes really is quite nonuniform. Here n is the ratio of
length to diameter. The demagnetizing field is the reason why for centuries magnets
were condemned to take awkward shapes of bars or horseshoes to avoid substantial
self-demagnetization and why the most successful electromagnetic machines of
the nineteenth century were built around electromagnets rather than permanent
magnets. The hardened steel magnets of the day showed little coercivity and were
easily demagnetized. Demagnetizing fields are also the cause of ferromagnetic
domains. The shape constraint on permanent magnets was not lifted until the middle
of the twentieth century. Permanent magnets then came to the fore in the design of
electric motors and magnetic devices. Fig. 6 illustrates a collection of magnets from
the eighteenth, nineteenth, and twentieth centuries.
The imaginative world of Maxwell and his followers in the latter part of the
nineteenth century when the electromagnetic revolution was in full swing was
16 J. M. D. Coey

Fig. 6 Magnets from four centuries; top, seventeenth-century lodestone, nineteenth-century elec-
tromagnet; bottom, eighteenth-century horseshoe magnet, twentieth-century alnico and Nd2 Fe14 B
magnets (not to scale)

actually far removed from our own [15]. They envisaged light and other Hertzian
waves as propagating in an all-pervasive aether, which was believed to possess
magical mechanical properties – it had to be a massless incompressible fluid,
transparent and devoid of viscosity, yet millions of times more rigid than steel!
Elaborate mechanical models were envisaged for the waves and fields. In due course
it came to be understood that reality was represented by the abstract mathematics,
which remained after all the mechanical props had been discarded.

The Earth’s Magnetic Field

The Earth’s field was the prime focus of attention of magnetism for over a millen-
nium, especially after it was understood that the magnetic field was of terrestrial
origin. By the beginning of the nineteenth century, the components of the field were
1 History of Magnetism and Basic Concepts 17

being recorded regularly in laboratories across the world. A comparison of the daily
magnetic records at Paris and Kazan, cities lying 4000 km apart, for the same day
in 1825, showed astonishingly similar short-term fluctuations. This inspired Carl
Friedrich Gauss to establish a worldwide network of 50 magnetic observatories,
coordinated from Göttingen, to make meticulous simultaneous measurements of the
Earth’s field, in the hope that if enough high-quality data could be collected, the
mystery of its origin and its fluctuations might be solved. This heroic pioneering
venture in international scientific collaboration amassed stores of data that were
enormous for that time. It inspired Gauss to develop spherical harmonic analysis,
from which he calculated that the leading, dipolar term accounted for about 90%
of the field and that the origin of the stable component was essentially internal.
Edward Sabine later spotted that the intensity of the short-term fluctuations tracked
the 11-year sunspot cycle, which we now know corresponds to reversals of the
solar magnetic field. But in its primary aim, Gauss’s Magnetische Verein must be
counted a failure. No amount of data, however copious and precise, could reveal a
deterministic origin of a phenomenon that was fundamentally chaotic. Piles of data
with no theory or hypothesis through which to view and be tested by them are not
very informative. This lesson was learned slowly.
The pole picture of the Earth’s magnetic field, albeit with poles that needed to
travel tens of kilometers every year to account for the secular variation, yielded
eventually in the academy if not in the popular imagination to one based on
electric currents driven by convection in the Earth’s liquid core. Joseph Larmor,
a dogged believer in the aether, was an early proponent of the geomagnetic dynamo.
He demonstrated the precession of a magnet in a magnetic field at a frequency
fL = γB/2π that bears his name. The precession is analogous to that of a spinning
top in a gravitational field; it is a consequence of the torque on a magnetic moment
expressed by Eq. 3. The constant γ, known as the gyromagnetic ratio, is the ratio of
the magnetic moment to its associated angular momentum. The proportionality of
these two quantities that at first sight appear quite dissimilar, the famous Einstein-de
Haas effect, was eventually demonstrated experimentally in 1915 (Fig. 7).

Fig. 7 The Einstein-de Haas


experiment. The iron rod
suspended from a torsion
fiber twists when a
magnetizing current in the
surrounding solenoid is
reversed, thereby
demonstrating the
relationship between
magnetism and angular
momentum
18 J. M. D. Coey

The Properties of Ferromagnets

If the luminiferous aether was inaccessible to experimental investigation, as the


1887 Michelson-Morley experiment suggested, the same could not be said for
magnetic materials. With its focus on electromagnetism, the nineteenth century
brought a flurry of investigations of the magnetic properties of the ferromagnetic
metals, iron (discovered in the fourth millennium BC), cobalt (discovered in 1735),
and nickel (discovered in 1824) and some of their alloys, which were at the
heart of electromagnetic machines. In 1842 James Joule, a brewer and natural
philosopher, discovered the elongation of an iron bar when it was magnetized
to saturation and demonstrated in a liquid displacement experiment that the net
volume was unchanged in the magnetostrictive process, owing to a compensating
contraction in the perpendicular directions [16]. Magnetostriction is the reason why
transformers hum. Gustav Wiedemann observed that an iron bar twisted slightly
when a current was passed through it in the presence of a magnetic field. Anisotropic
magnetoresistance (AMR) was discovered by William Thomson in 1856; the
resistance of iron or nickel is a few percent higher when measured in the direction
parallel to the magnetization than in the perpendicular direction [17]. The Hall
effect, the appearance of a transverse voltage when a current was passed through
a gold foil subject to a transverse magnetic field was discovered by Edwin Hall in
1879, And the contribution e proportioal to the magnetization of a ferromagnet —
tha anomalous Hall effect — was found shortly afterwards, in iron. John Kerr
showed in 1877 that the rotation of the plane of polarization of electromagnetic
radiation, demonstrated by Faraday for light passing through glass, could also be
measured in reflection from polished ferromagnetic metal surfaces [18].
Gauss’s collaborator Wilhelm Weber, who had constructed the first electromag-
netic telegraph in 1833, formally presented the idea that molecules of iron were
capable of movement around their centers, suggesting that they lay in different
directions in an unmagnetized material, but aligned in the same direction in the
presence of an applied magnetic field. This was the origin of the explanation of
hysteresis by James Alfred Ewing, who coined the name for the central phenomenon
of ferromagnetism that he illustrated using a board of small, pivoting magnets [19].
Ewing’s activities as a youthful scottish professor at the University of Tokyo in
the 1890s helped to establish the strong Japanese school of research on magnetic
materials that thrives to the present day.
The hysteresis loop, illustrated in Fig. 8, is the icon of ferromagnetism. Except
in very small particles, a magnetized state is always metastable. The saturated
magnetic state is higher in energy relative to a multidomain state on account of the
demagnetizing field that creates a positive magnetostatic self-energy -½μ0 Ms .Hd
dV in the fully magnetized state, where the only contribution to the integral comes
from the magnet volume. The hardened steel magnets of the nineteenth century
showed little coercivity, Hc  Ms , and could only survive as bars and horseshoes
where the demagnetizing factor N of Eq. 11 was 1. The principal achievement
in technical magnetism in the twentieth century was the mastery of coercivity; this
needed new materials having Hc  Ms .
1 History of Magnetism and Basic Concepts 19

M
spontaneous magnetization

remanence

coercivity virgin curve


initial susceptibility H

major loop

Fig. 8 The hysteresis loop of magnetization M against magnetic field H for a typical permanent
magnet, showing the initial magnetization curve from the equilibrium multidomain state and the
major loop. Ms is the saturation magnetization, Mr the remanent magnetization at zero field, and
Hc the coercive field required to reduce the magnetization to zero

The astonishing transformation of science and society that began in 1820


deserves the name electromagnetic revolution. By the end of the century, elec-
tromagnetic engineering was electrifying the planet, changing fundamentally our
communications and the conditions of human life and leisure. Huge electric
generators, powered by hydro or fossil fuel, connected to complex distribu-
tion networks were bringing electric power to masses of homes and factories
across the Earth. Electric light banished the tyranny of night. Electric motors
of all sorts were becoming commonplace, and public transport was transformed.
Telegraph and telephone communication connected people across cities, coun-
tries, and continents. Valdemar Poulsen demonstrated magnetic voice recording
in 1898. Much of the progress was achieved by engineers who relied on prac-
tical knowledge of electrical circuits and magnetic materials, independently of
the conceptual framework of electrodynamics that had been developed by the
physicists.
The electromagnetic revolution and the subsequent electrification of the planet
were the second occasion when magnetism changed the world. The century closed
with Pierre Curie’s 1895 accurate measurements of the Curie point TC (the critical
temperature above which a material abruptly loses its ferromagnetism) and with the
all-important discovery of the electron. Yet ferromagnetism was hardly understood
at all at a fundamental level at the turn of the century, and it was becoming evident
that classical physics was not up to the task.
20 J. M. D. Coey

Magnetism of the Electron

The discovery of the electron in the closing years of the nineteenth century was a
huge step toward the modern understanding of magnetism. The elementary charged
particle with mass me = 9.109 10−31 kg and charge e = −1.602 10−19 C had been
named by the Irish scientist George Johnstone Stoney in 1891, several years before
Jean Perrin in France actually identified negatively charged particles in a cathode
ray tube and J. J. Thompson in England measured their charge to mass ratio e/me ,
by deflecting the electrons in a magnetic field and making use of Eq. 6. Another
Irish scientist, George Francis FitzGerald, suggested in 1900 that magnetism might
be due to rotational motion of these electrons. They turned out to be not only the
carriers of electric current but also the essential magnetic constituent of atoms and
solids.

The Demise of Classical Physics

At the beginning of the twentieth century, the contradictions inherent in contempo-


rary physics could no longer be ignored, but 25 years were to elapse before they
could be resolved. In that heroic period, classical physics and the lingering wisps
of aether were blown away, and a new paradigm was established, based on the
principles of quantum mechanics and relativity.
Magnetism in particular posed some serious puzzles. In order to account for the
abrupt disappearance of ferromagnetism at the Curie point, Pierre Weiss, who had
developed Ewing’s concept of magnetic domains, postulated in 1907 the existence
of an internal molecular field.

H i = nW M (12)

proportional to magnetization in order to explain the spontaneous magnetization


within them. His theory of ferromagnetism was based on Paul Langevin’s 1905
explanation of the Curie law susceptibility of an array of disordered classical
magnetic moments.

χ = C/T (13)

Susceptibility χ can be conveniently defined as the dimensionless ratio M/H,


where H is the applied magnetic field. The expression is modified for a ferromagnet
above its Curie point where it becomes the Curie-Weiss law χ = C/(T – θp ) with
θp ≈ TC . With Eq. (12) and Langevin’s theory of paramagnetism, Weiss invented
the first mean-field theory of a phase transition. For iron, where M = 1.71 MAm−1 ,
the Weiss constant nW is roughly 1000. According to Maxwell’s equation ∇. B = 0,
the component of B normal to the surface of a magnet is continuous, so there should
1 History of Magnetism and Basic Concepts 21

be a stray field of order μ0 Hs ∼ 1000 T in the vicinity of a magnetized iron bar. In


fact, the observed stray fields are a thousand times smaller.
Furthermore if, as Ampère believed, all magnetism was traceable to circulating
electric currents, the magnetization of an iron bar requires an incredible surface
current of 17,100 A for every centimeter of its length. How could such a current
be sustained indefinitely? Why does the iron not melt? What did the sobriquet
molecular really mean? The anomalous Zeeman splitting of spectral lines in a
magnetic field was another mystery. In retrospect, the most startling result was a
theorem proved independently in their theses by Niels Bohr in 1911 and Hendrika
van Leeuwen in 1919. They showed that at any finite temperature and in any
magnetic or electric field, the net magnetization of a collection of classical electrons
vanishes identically. So, in stark contrast with experiment, classical electron physics
was fundamentally incompatible with any kind of magnetism!
By 1930, quantum mechanics and relativity had ridden to the rescue, and a
new understanding of magnetism emerged in terms of the physics of Einstein,
Bohr, Pauli, Dirac, Schrödinger, and Heisenberg. The source of magnetism in
matter was identified with the angular momentum of elementary particles, especially
the electron [20]. The connection between angular momentum and magnetism
had been demonstrated directly on a macroscopic scale in 1915 by the Einstein-
de Haas experiment (Fig. 7), where angular recoil of a suspended iron rod was
observed when its magnetization was reversed by an applied field. It turned out
that the perpetual currents in atoms were quantized in stationary states that did not
decay and that the angular momentum of the orbiting electrons was a multiple of
Planck’s constant  = 1.055 10−34 Js. Furthermore, the electron itself possessed an
intrinsic angular momentum or spin [20] with eigenvalues of ±½ along the axis of
quantization defined by an external field. Weiss’s molecular field was no magnetic
field at all, but a manifestation of electrostatic coulomb interactions constrained by
Wolfgang Pauli’s exclusion principle, which forbade the occupancy of a quantum
state by two electrons with the same spin.
The intrinsic angular momentum of an electron with two eigenvalues had been
proposed by Pauli in 1924; Samuel Goudsmit and George Uhlenbeck demonstrated
a year later that the spin angular momentum had a value of ½. The Pauli spin
matrices representing the three components of spin angular momentum are




01 0 −i 1 0
s= , , /2 (14)
10 i 0 0 −1
The corresponding electronic magnetic moment was the Bohr magneton,

μB = e/2me

or 9.274 × 10−24 Am2 , twice as large as the moment associated with a unit of
orbital angular momentum in Bohr’s model of the atom. The gyromagnetic ratio
of magnetic moment to angular momentum for the electron spin is γ ≈ e/me ,
so the Larmor precession frequency eB/2πme for the electron is 28 GHzT−1 .
22 J. M. D. Coey

The problem of the electron’s magnetism was finally resolved by Paul Dirac
in 1928 when he succeeded in writing Schrödinger’s equation in relativistically
invariant form, obtaining the non-relativistic electron spin in terms of the 2 × 2
Pauli matrices. Together with Dirac, Werner Heisenberg formulated the exchange
interaction represented by the famous Heisenberg Hamiltonian

H = –2J S i .S j (15)

to describe the coupling between the vector spins Si and Sj of two nearby many-
electron atoms i and j. The spin vectors S are the spin angular momenta in units of .
The value of the exchange integral J was closely related to Weiss’s molecular field
coefficient nW and depends strongly on interatomic distance. It can be positive, if it
tends to align the two spins parallel (ferromagnetic exchange), or negative if it tends
to align the pair antiparallel (antiferromagnetic exchange). The value of S is obtained
from the first of the three rules, discussed below, that were formulated by Friedrich
Hund around 1927 for finding the ground state of a multi-electron atom. The
exchange interactions among the electrons of the same atom are much stronger than
those between the electrons of adjacent atoms given by Eq. (15). The fundamental
insight that magnetic coupling of electronic spins is governed by electrostatic
coulomb interactions, subject to the symmetry constraints of quantum mechanics,
was the key needed to unlock the mysteries of ferromagnetism. Exchange is
discussed in  Chap. 2, “Magnetic Exchange Interactions.”
The magnetic moment of an atom or ion is the sum of two contributions. One
arises from the intrinsic spin angular momentum of the atomic electrons. The other
comes from their quantized orbital angular momentum. The moments associated
with each type of angular momentum have to be summed according to the rules of
quantum mechanics. The moment associated with ½ of spin angular momentum is
practically identical to that associated with  of orbital angular momentum, namely,
one Bohr magneton in each case. The quantum theory of magnetism is therefore the
quantum theory of angular momentum. Hund’s rules were an empirical prescription
for determining the total angular momentum of the many-electron ground state of
electrons belonging to the same atom or ion. Firstly, the rule is to maximize the
spin angular momentum S while respecting the Pauli principle that no two electrons
can be in the same quantum state. Secondly, the orbital angular momentum L is
maximized, consistent with the value of S, and thirdly the spin and orbital momenta
are coupled together to form the total angular momentum J = L ± S, according
to whether the electronic shell is more or less than half full. The total magnetic
moment (in units of μB ) is then related to the total angular momentum (in units of
) by a numerical Landé g-factor, which is 1 for a purely orbital moment and 2 for
pure spin.
The spin-orbit coupling, which arises in the atom from motion of the electron
in the electrostatic potential of the charged nucleus and gives rise to Hund’s third
rule, is another key interaction. Of fundamentally relativistic character, it emerges
naturally from Dirac’s relativistic quantum theory of the electron, and it turns out to
be at the root of many of the most interesting phenomena in magnetism, including
1 History of Magnetism and Basic Concepts 23

magneto-optics, magnetocrystalline anisotropy, and the spin Hall effect. The spin-
orbit interaction for a magnetic ion is represented by the Hamiltonian L.S, where
L is the orbital angular momentum of the many-electron atom in units of  and 
is the atomic spin-orbit coupling constant. Like the exchange constant J ,  has
dimensions of energy.
Felix Bloch in 1930 described the spin waves that are the quantized elementary
excitations of a ferromagnetic array of atoms whose spins are coupled by Heisen-
berg exchange. These excitations have an angular frequency ω and a wavevector
k that are related by the dispersion relation ω = Dk2 , where D is the spin wave
stiffness constant. It is proportional to J .
The first quantum theories of magnetism regarded the electrons as localized
on the atoms or ions, but an alternative magnetic band theory of ferromagnetic
metals was developed by John Slater and Edmund Stoner in the 1930s. It accounted
for the non-integral, delocalized spin moments found in Fe, Co, and Ni and their
alloys, although the theory in its original form greatly overestimated the Curie
temperatures. The delocalized, band electron model of Slater and the localized,
atomic electron model of Heisenberg were two distinct paradigms for the theory
of magnetism that persisted until sophisticated computational methods for treating
the many-body interelectronic correlations in the ground state of multi-electron
atoms were devised toward the end of the twentieth century. The differences
between the two approaches are epitomized in the calculation of the paramagnetic
susceptibility. Pauli found a small temperature-independent susceptibility resulting
from Fermi-Dirac statistics for delocalized electrons, whereas Léon Brillouin had
used Boltzmann statistics and the Bohr model to derive the Curie law susceptibility
of an array of atoms with localized electrons.
The sixth Solvay Conference, held in Brussels in October 1930 (Fig. 9), was
devoted to magnetism [21]. It followed four years of brilliant discoveries in
theoretical physics, which set out the modern electronic theory of condensed matter.
Yet the immediate impact on the practical development of functional magnetic
materials was surprisingly slight. Dirac there made the perceptive remark “The
underlying physical laws necessary for the mathematical theory of a large part of
physics and the whole of chemistry are completely known, and the difficulty is only
that the exact application of these laws leads to equations much too complicated to
be soluble.”

Magnetic Phenomenology

In view of the immense computational challenge posed by many-body electron


physics in 1930, a less fundamental theoretical approach was needed. Louis
Néel pursued a phenomenological approach to magnetism with notable success,
oblivious to the triumphs of quantum mechanics. His extension of the Weiss
theory to two equal but oppositely aligned magnetic sublattices led him to the
idea of antiferromagnetism in his 1932 doctoral thesis. This hidden magnetic order
24 J. M. D. Coey

Fig. 9 The 1930 Solvay Conference on Magnetism Back row: Herzen, Henriot, Verschaffelt,
Manneback, Cotton, Errera, Stern, Piccard, Gerlach, Darwin, Dirac, Bauer, Kapitza, Brioullin,
Kramers, Debye, Pauli, Dorfman, van Vleck, Fermi, Heisenberg. Front row: de Donder, Zeeman,
Weiss, Sommerfeld, Curie, Langevin, Einstein, Richardson, Cabrera, Bohr, de Haas

awaited the development of neutron scattering in the 1950s before it could be


directly revealed, initially for MnO. Néel went on to explain the ferrimagnetism
of oxides such as magnetite, Fe3 O4 , the main constituent of lodestone, in terms of
two unequal, antiferromagnetically coupled sublattices. The three most common
types of magnetic order, and their temperature dependences, are illustrated in
Fig. 10.
The spinel (MgAl2 O4 ) structure of magnetite has an A sublattice of 8a sites with
fourfold tetrahedral oxygen coordination and twice as many 16d sites with sixfold
octahedral coordination forming a B sublattice. The spinel structure is illustrated
in Fig. 14 where the 8a sites are at the centers of the blue tetrahedra, which have
oxygen ions at the four corners, and the 16d sites are at the centers of the brown
octahedra, which have six oxygen ions at the corners. The numbers of each type
of site in the unit cell are indicated by the labels. The 16d sites in magnetite
are occupied by a mixture of ferrous Fe2+ and ferric Fe3+ ions with electronic
configurations 3d5 and 3d6 and spin moments of 5 μB and 4 μB , respectively,
whereas the 8a sites are occupied by oppositely aligned Fe3+ ions. This yields a net
spin moment of 4 μB per formula (0.48 MAm−1 ) – a quantitative explanation of the
magnetism of the archetypical magnet in terms of lattice geometry and the simple
rule that each unpaired electron contributes a spin moment of one Bohr magneton.
Néel added two new categories of magnetic substances – antiferromagnets and
ferrimagnets – to Faraday’s original three. Their magnetic ordering temperatures are
known as antiferromagnetic or ferrimagnetic Néel temperatures. The ferrimagnetic
one is also called a Curie point.
1 History of Magnetism and Basic Concepts 25

1/c 1/c 1/c

T C ,qp T qp TN T qp T fN T

M M M

B B
B A B
A

TC T TN T T fN T

A
A

Fig. 10 Schematic temperature dependences of the inverse susceptibility (top) and (sub)lattice
magnetization (bottom) of a ferromagnet (left), an antiferromagnet (center), and a ferrimagnet
(tight)

Micromagnetism

For many practical purposes, it is possible to follow in the footsteps of Néel,


sidestepping the complications engendered by the atomic and electronic basis of
magnetism, and regard magnetization as a continuous vector in a solid continuum
[13], as people have for about 200 years. The iconic hysteresis loop M(H) (Fig. 8)
is the outcome of a metastable structure of domains of uniformly magnetized
ferromagnetic Weiss domains separated by narrow domain walls between domains
magnetized in different directions. The structure depends on the thermal and
magnetic history of a particular sample. Aural evidence for discontinuous jumps
in the size of the domains as the magnetization was saturated was first heard by
Heinrich Barkhausen in 1919 with the help of a pickup coil wound around some
ferromagnetic wires, a rudimentary amplifier, and a loudspeaker. Then in 1931 the
domains were directly visualized by Francis Bitter using a microscope focused on a
polished sample surface and a colloidal suspension of magnetite particles that were
drawn by the stray field to the domain walls. These colloids, known as ferrofluids,
behave like ferromagnetic liquids.
The idea of a domain wall as a region where the magnetization rotates pro-
gressively from one direction to the opposite one in planes parallel to the wall
was introduced by Felix Bloch in 1932. His walls create no bulk demagnetizing
26 J. M. D. Coey

Fig. 11 Two types of 180◦ domain walls: a) the Bloch wall and b) the Néel wall

field and cost little magnetostatic energy because ∇. M = 0; the magnetization


in each plane is uniform, and there is no component perpendicular to the planes
(see Eq. 9). The exchange energy cost, written in the continuum approximation as
A(∇M)2 where A ∝ J , is balanced by the anisotropy energy cost associated with the
magnetization in the wall that is misaligned with respect to a magnetic easy axis of
the crystal. Magnetic anisotropy is introduced below, and it is discussed in detail in
 Chap. 3, “Anisotropy and Crystal Field.” A Néel domain wall, where the
magnetization rotates in a plane perpendicular to the wall so that ∇. M = 0 in the
bulk, but there is no surface magnetic charge, is higher in energy except in thin films.
The two types of wall are illustrated in Fig. 11.
In principle, the sum of free energy terms associated with exchange, anisotropy,
and magnetostatic interactions, together with the Zeeman energy in an external field,
could be minimized to yield the M(H) loop and the overall domain structure of
any solid. Further terms can be added to take into account the effects of imposed
strain and spontaneous magnetostriction. In practice, however, crystal defects such
as grain boundaries spoil the continuum picture and can exert a crucial influence on
the walls. It is then necessary to resort to models to develop an understanding of
hysteresis.
The basic theory of micromagnetism was developed by William Fuller Brown
in 1940 [13]. The magnetostatic interaction between the magnetic dipoles that
constitute the magnetization is a dominant factor. The dipole fields fall off as 1/r3
(Eq. 2), providing a long-range interaction unlike exchange, which is short-range
because it depends on an overlap between electronic wavefunctions that decays
exponentially with interatomic spacing. This is why weak magnetostatic interactions
that are of order 1 K for a pair of ions are able to compete on a mesoscopic length
scale with the much stronger exchange interactions of electrostatic origin that can
be of order 100 K to control the domain structure of a given ferromagnetic sample.
Magnetocrystalline anisotropy is represented phenomenologically in the theory
by terms in the energy that depend on the orientation of M with respect to the
local crystal axes. The electrostatic interaction of localized atomic electrons with
the potential created by all the other atoms in the crystal is known as the crystal field
1 History of Magnetism and Basic Concepts 27

interaction; the effect of chemical bonding with the ligands of an atom is the ligand
field interaction. The two effects are comparable in magnitude for 3d ions [22].
Magnetocrystalline anisotropy arises from the interplay of the crystal/ligand field
and spin-orbit coupling. The simplest case is for uniaxial (tetragonal, hexagonal,
rhombohedral) crystals, where the leading term in the energy density is of the form

Ea = K1 sin2 θ + . . . .. (16)

where θ is the angle between M and the symmetry axis. Two opposite easy
directions lie along the crystal axis if the anisotropy constant K1 is positive, but
there are many easy directions lying in an easy plane perpendicular to the crystal
axis (θ = π/2) when K1 is negative. Anisotropy arises also from overall sample
shape, due to the demagnetizing energy ½MHd , which gives another contribution in
sin2 θ that depends on the demagnetizing factor N with

1
K1 sh = μ0 Ms 2 (1 − 3N ) (17)
4

where Ms is the spontaneous magnetization. There is obviously no shape anisotropy


for a sphere, which has N = 1/3. An expression equivalent to (16) at the atomic scale
is εa = Da sin2 θ , where Da /kB ∼ 1 K. The magnitude of the crystal field energy is
comparable to the magnetostatic energy, but it is much smaller than the exchange
energy in practical magnetic materials. It remains challenging to calculate K1 or Da
precisely in metals.
An instructive paradox arising from Brown’s micromagnetic theory is his result
that the coercivity Hc of a perfect, defect-free ferromagnetic crystal lattice must
exceed the anisotropy field Ha = 2 K1 /μ0 Ms . In practice Hc is rarely as much as a
fifth of Ha . The explanation is that no real lattice is ever free of defects, which act as
sites for the nucleation of reverse domains or as pinning centers for domain walls.
The sequence of metastable states represented on the hysteresis loop is generally
dominated by asperities and lattice defects that are very challenging to characterize
in any real macroscopic sample. Control of these defects in modern permanent
magnets having Hc  Ms has been as much a triumph of metallurgical art as physical
theory. Micromagnetism is the subject of  Chap. 7, “Micromagnetism.”

Magnetic Materials

The traditional magnetic materials were alloys of the ferromagnetic metals, Fe,
Co, and Ni. The metallurgy and magnetic properties of these alloy systems were
the focus of investigations of technical magnetism in the first half of the twentieth
century, when useful compositions were developed such as Permendur, Fe50 Co50 ,
the alloy with the highest magnetization (1.95 MAm−1 ); Permalloy Fe20 Ni80 , which
has near-zero anisotropy and magnetostriction, together with very high relative
permeability (μr = (1 + χ) ≈ 105 ); and Invar Fe64 Ni36 a composition with near-
28 J. M. D. Coey

Fig. 12 Unit cells of the ferromagnetic elements Fe (body-centered cubic, left), Ni (face-centered
cubic, center), and Co, Gd (hexagonal close-packed, right) [29], with kind permission from
Cambridge University Press

zero thermal expansion around room temperature. The early investigations are well
summarized in Bozorth’s 1950 monograph [23]. The fourth ferromagnetic element
at room temperature is the rare earth gadolinium. The crystal structures of these
elemental ferromagnets are illustrated in Fig. 12.
An important practical advance in the story of permanent magnet development
was the thermal processing of a series of Al-Ni-Co-Fe alloys, the Alnico magnets,
that was initiated in Japan in 1932 by Tokushichi Mishima. Their coercivity relied
on achieving a nanostructure of aligned acicular (needle-like) regions of Co-Fe in
a matrix of nonmagnetic Ni-Al. It was the shape of the ferromagnetic regions that
gave the alloys some built-in magnetic anisotropy (Eq. 17), but it still had to be
supplemented with global shape anisotropy by fabricating the Alnico into a bar or
horseshoe in order to avoid self-demagnetization. The mastery of coercivity that
was acquired over the course of the twentieth century (Fig. 13) was spectacular,
and burgeoning applications in technical magnetism of soft and hard magnetic
materials were the direct consequence. The terms “soft” and “hard” were derived
originally from the magnetic steels that were used in the nineteenth century. The
most useful figure of merit for the hard, permanent magnets is the maximum energy
product |BH|max , equal to twice the energy in the stray field produced by a unit
volume of magnet. The SI unit is kJm−3 . Energy product doubled every 12 years
for most of the twentieth century, thanks to the discovery in the 1960s of rare
earth cobalt intermetallic compounds and the discovery of new rare earth iron-
based materials in the 1980s. Comparable progress with decreasing hysteresis losses
in soft, electrical steels continued to the point where they became a negligible
fraction of the resistive losses in the copper windings of electromagnetic energy
converters. Ultrasoft amorphous magnetic glasses were developed in the 1970s.
Applications of soft and hard magnetic materials are discussed in  Chaps. 29, “Soft
Magnetic Materials and Applications,” and  28, “Permanent Magnet Materials and
Applications” respectively.
A good working knowledge of the quantum mechanics of multi-electron atoms
and ions had been developed by the middle of the twentieth century, mainly from
1 History of Magnetism and Basic Concepts 29

Nd-Fe-B

Fig. 13 The development of coercivity over the ages and in the twentieth century

observations of optical spectra and the empirical rules formulated by Hund to


specify the ground state L, S, and J multiplet, which is the one of interest for
magnetism. All this led naturally to a focus on the localized electron magnetism
found in the 3d and 4f series of the periodic table. For 3d ions in solids, the
ionic moment is essentially that arising from the unpaired electron spins left after
filling the orbitals according to the Pauli principle and Hund’s first rule. The orbital
moment expected from the second rule is quenched by the crystal field, which
impedes the orbital motion so that it barely contributes to the ionic magnetism.
But the crystal field is weaker for the 4f elements in solids, whether insulating or
metallic, and the magnetism is more atomic-like with spin and orbital contributions
coupled by the spin-orbit interaction according to Hund’s third rule to yield the total
angular momentum J.
Microscopic quantum theory began to play a more important part in magnetic
materials development after the 1970s with the advent of rare earth permanent
magnets SmCo5 and especially Nd2 Fe14 B, when an understanding of the intrinsic,
magnetocrystalline anisotropy in terms of crystal field theory and spin-orbit cou-
pling began at last to make a contribution to the design of new permanent magnet
materials.

Magnetic Oxides

The focus on localized electron magnetism in the 1950s and 1960s led to systematic
investigations of exchange interactions in insulating compounds where the spin
30 J. M. D. Coey

moments of magnetic 3d ions are coupled by indirect overlap of their wavefunc-


tions via an intervening nonmagnetic anion, usually O2− . A systematic empirical
understanding of the dependence of these superexchange interactions on electron
occupancy and bond angle emerged in the work of Junjiro Kanamori and John
Goodenough [24], based on the many new magnetic compounds that were being
fabricated at that time. There is a multitude of solid solutions between end-members,
with extensive opportunities to tune magnetic properties by varying the chemical
compositions of oxide families such as ferrites [25]. Superexchange, like direct
exchange in the ferromagnetic 3d elements, depends on the overlap of wavefunc-
tions of adjacent atoms and decays exponentially with interatomic distance.
The magnetite family of cubic spinel ferrites M2+ Fe3+ 2 O4 was the first to be
thoroughly investigated, with M = Mg, Zn, Mn, Fe, 2/3Fe3+ (γFe2 O3 ), Co, or Ni.
Ferrimagnetic Neél temperatures of these ferrites range from 700 to 950 K, although
spinel itself (MgAl2 O4 ) is nonmagnetic. Several of the insulating compounds
with Mn, Ni, and Zn are suitable as soft magnetic materials for audio- or radio-
frequency applications. Other important families investigated at that time were
garnets, perovskites, and hexagonal ferrites. The garnet ferrites R3+ 3+
3 Fe5 O12 have
a large cubic unit cell containing 160 ions, with ferrimagnetically aligned ferric
iron in both tetrahedral 24d and octahedral 16a sites, and large R3+ ions in
eightfold oxygen coordination in deformed cubal 24c sites. R may be any rare
earth element, including Y, which forms yttrium iron garnet (YIG), Y3 Fe5 O12 , a
superlative microwave material that exhibits ultra-low magnetic losses on account
of its insulating character. The net magnetic moment of YIG is 5μB per formula
unit. Substituting magnetic rare earths in the structure provides an opportunity to
study superexchange between 3d and 4f ions. That interaction is weak, and the
4f ions couple antiparallel to the 24d site iron, but their sublattice magnetization
decays much faster with temperature, giving rise to the possibility of a compensation
temperature, where the net magnetization of the two ferrimagnetic sublattices
crosses zero at a temperature below the ferrimagnetic Neél point. The compensation
temperature of Gd3 Fe5 O12 , for example, is 290 K, whereas its ferrimagnetic Néel
point is at 560 K, a typical value for the whole rare earth iron garnet series.
Another important oxide family, the hexagonal ferrites especially M2 Fe12 O19 ,
where M = Ba2+ or Sr2+ , have uniaxial anisotropy and crystallize in the magneto-
plumbite structure. There are four Fe3+ sites in the structure, including a fivefold 2b
site with trigonal symmetry where the threefold axis is parallel to the c-axis of the
hexagonal unit cell. The net ferrimagnetic moment is 20 μB per formula unit, since
eight iron ions belong to one sublattice and four to the other. The large nonmagnetic
M cations occupy sites that would otherwise belong to a hexagonal close-packed
oxygen lattice. The 2b site contributes rather strong uniaxial anisotropy, and the
anisotropy field of 1.4 MAm−1 is more than three times the magnetization (0.38
MAm−1 ), making it possible in the early 1950s to achieve coercivity comparable
to the magnetization and manufacture cheap ceramic magnets in any desired shape,
thereby overcoming the shape barrier that had impeded the development permanent
magnets for a millennium. A million tonnes of these ferrite magnets is sold every
year.
1 History of Magnetism and Basic Concepts 31

The drawback of any oxide magnetic material is that its magnetization is never
more than a third of that of metallic iron. This is unavoidable because most of the
unit cell volume is occupied by large, nonmagnetic O2− anions, with the high-spin
ferric iron Fe3+ or other magnetic ions confined to the interstices in the oxygen
lattice. To make matters worse, a ferrimagnetic structure reduces the magnetization
further. There are relatively few ferromagnetic oxides; CrO2 is one example. It is
not an insulator, but a half metal, with a gap in the minority-spin conduction band.
A search for insulating ferromagnetic oxides in the 1950s led to the investi-
gation of ABO3 compounds with the perovskite structure. Here the magnetic B
cations occupy the 1a octahedral sites, and the nonmagnetic A cations occupy
the 12-coordinated 1b sites in the ideal cubic structure. It proved to be possible
to obtain ferromagnetism provided the A cations are present in two different
valence states. This works best in mixed-valence manganites [26], with composition
(La3+ 2+
0.7 M0.3 )MnO3 where M = Ba, Ca, or Sr. The resulting mixture of Mn
3+ (3d4 )
4+ 3
and Mn (3d ) on B sites leads to electron hopping with spin memory from
one 3d3 core to another. This is the ferromagnetic double exchange interaction,
envisaged by Clarence Zener in 1951. Similar electron hopping occurs for Fe2+
and Fe3+ in the octahedral sites of magnetite. A consequence is that the oxides,
though ferromagnetic, are no longer insulating, and the Curie temperatures are not
particularly high – they do not exceed 400 K. A notable feature of the mixed-valence
manganites, related to their hopping conduction, is the “colossal magnetoresistance”
observed near the Curie point, where there is a broad maximum in the resistance
that can be suppressed by applying a magnetic field of several tesla. All four
oxide structures are presented in Fig. 14. They illustrate the importance of crystal
chemistry for determining magnetic properties.

Fig. 14 Crystal structures of magnetic oxides: perovskite (top left), spinel (bottom left), garnet
(center), magnetoplumbite (right). The oxygen coordination polyhedral around the magnetic
cations (tetrahedrons, blue, or octahedrons, brown) is illustrated. The spheres are large nonmag-
netic cations. Unit cells are outlined in black. Magnetoplumbite is hexagonal, and the others are
cubic [31], with kind permission from APS
32 J. M. D. Coey

Research on localized electron magnetism in oxides and related compounds has


passed through three phases. Beginning with studies of polycrystalline ceramics
from about 1950, single crystals were grown for specific physical investigations
after about 1970, and then in the late 1980s, following the high-temperature
superconductivity boom, came the growth and characterization of ferromagnetic
and ferrimagnetic oxide thin films and first steps toward all-oxide spin electronics. A
similar pattern was followed by sulfides, fluorides, and other magnetic compounds.
All are discussed further in  Chap. 17, “Magnetic Oxides and Other Compounds.”

Intermetallic Compounds

A rich class of functional magnetic materials is the intermetallic compounds of


rare earth elements and transition metals. The atomic volume ratio of a 4f to a
3d atom is about three, so the alloys tend to be stoichiometric line compounds
rather than solid solutions. The first of these was SmCo5 , developed for permanent
magnet applications in the USA in the mid-1960s by Karl Strnat. It was followed
by Sm2 Co17 in the early 1970s, and then in 1983 came the announcement of
the independent discovery of the first iron-based rare earth magnet, the ternary
Nd2 Fe14 B, by Masato Sagawa in Japan and John Croat in the USA. This was
a breakthrough because iron is cheaper and more strongly magnetic than cobalt.
Nd2 Fe14 B has since come to dominate the global high-performance magnet market,
with an annual production in excess of 100,000 tonnes. The coercivity needed in
these optimized rare earth permanent magnets is comparable to their magnetization,
and the optimization of the microstructure of a new hard magnetic material to attain
the highest possible energy product, which scales as Ms 2 but can never exceed
¼μ0 Ms 2 , is a long empirical process. It generally takes many years to achieve
a coercivity as high as 20–30% of the anisotropy field [28]. The battle to create
the metastable hysteretic state that permits a permanent magnet to energize the
surrounding space with a large stray field is never easy to win, and each material
requires a different strategy.
The fundamental significance of these intermetallics and related interstitial
compounds such as Sm2 Fe17 N3 that were discovered in the 1990s is that crystal field
theory and quantum mechanics were involved in their design. All have a uniaxial
crystal structure with a single easy axis and strong magnetocrystalline anisotropy.
Such anisotropy is a prerequisite for the substantial coercivity, Hc  Ms needed to
overcome the shape barrier and create a magnet with any desired form.
The practical significance of the rare earth permanent magnets has been the
appearance of a wide range of compact, energy-efficient electromagnetic energy
converters that are being used in consumer products, electric vehicles, aeronautics,
robotics, and wind generators.
Besides magnetocrystalline anisotropy, another potentially useful consequence
of the spin-orbit interaction in rare earth intermetallics is their strong magnetostric-
tion. The rare earth elements order magnetically at or below room temperature
so, just as for the permanent magnets, it was necessary to form an intermetallic
1 History of Magnetism and Basic Concepts 33

Fig. 15 Crystal structures of ferromagnetic intermetallic compounds: YFe2 (cubic, left) SmCo5
(hexagonal, top centre), Co2 MnSi (cubic, bottom centre), Nd2 Fe14 B (tetragonal, right). Fe and Co
Mn are the small brown/red, blue, and scarlet spheres. Rare earths are the large spheres. Si and B
are grey and black

compound with iron or cobalt to obtain a functional material with a useful


Curie temperature that should be substantially greater than room temperature to
ensure adequate magnetic stability. A functional magnetostrictive material has to
be magnetically soft, and this was achieved in the RFe2 rare earth Laves phase
compounds by Arthur Clark in 1984, who combined Dy and Tb, which have the
same sign of magnetostriction, but compensating anisotropy of opposite sign, in
the cubic alloy (Tb0.3 Dy0.7 )Fe2 , known as Terfenol-D. Single crystals exhibited
Joulian magnetostriction of up to 2000 parts per million (ppm), a hundred times
greater than Joule had measured 150 years earlier in pure iron [16] (see  Chaps.
28, “Permanent Magnet Materials and Applications,” and  11, “Magnetostriction
and Magnetoelasticity”).
Magnetically soft rare earth intermetallics are also of interest as magnetocaloric
materials for solid-state refrigeration when their Curie point is close to room
temperature (see  Chap. 30, “Magnetocaloric Materials and Applications”). Some
crystal structures of rare earth intermetallics are shown in Fig. 15.
Among the other intermetallic families, the ordered body-centered cubic Heusler
families of X2 YZ or XYZ alloys are notable in that they include a wide variety
of magnetically ordered compounds, such as the magnetic shape-memory alloy
NiMnSb or the half-metallic ferromagnet Co2 MnSi, which, like CrO2 , has a gap at
the Fermi level for minority-spin electrons. Information on a great many metallic
magnetic materials is collected in  Chap. 4, “Electronic Structure: Metals and
Insulators.”
34 J. M. D. Coey

Model Systems

Magnetism has proved to be a fertile proving ground for condensed matter theory.
The first mean-field theory was Weiss’s molecular field of magnetism, later
generalized by Lev Landau in the USSR in 1937. There followed more sophisticated
theories of phase transitions, with magnetism providing much of the data to support
them. The single-ion anisotropy of rare earth ions due to the local crystal field
reduces the effective dimensionality of the magnetic order parameter from three
to two for easy-plane (xy) anisotropy or from three to one for easy-z-axis (Ising)
anisotropy. Magnetically ordered compounds can be synthesized with an effective
spatial dimension of one (chains of magnetic atoms), two (planes of magnetic
atoms), or three (networks of magnetic atoms), as well as ladders and isolated
motifs. Magnetism has provided a treasury of materials that show continuous
phase transitions as a function of temperature or quantum phase transitions at zero
temperature as a function of pressure or magnetic field, as well as topological phases
such as the two-dimensional xy model, investigated by David Thouless, Michael
Kosterlitz, and Duncan Haldane. It is frequently possible to realize magnetic
materials that embody the essential electronic or structural features of the theoretical
models.
An early theoretical milestone was Lars Onsager’s 1944 solution of the two-
dimensional Ising model, where spins are regarded as one-dimensional scalars
that can take only values of ±1. The behavior of more complex and realistic
systems such as the three-dimensional Heisenberg model near its Curie temperature
was solved numerically using the renormalization group technique developed by
Kenneth Wilson in the 1970s. The ability to tailor model magnetic systems, with
an effective spatial dimension of 1 or 2 due to their structures of chains or planes
of magnetic ions and an effective spin dimension of 1, 2, or 3 determined by
magnetocrystalline anisotropy due to the combination of the crystal/ligand field and
the spin-orbit interaction, was instrumental in laying the foundation of the modern
theory of phase transitions. The theory is based on universality classes where
power-law temperature variations of the order parameter and its thermodynamic
derivatives with respect to temperature or magnetic field in the vicinity of the phase
transition are characterized by numerical critical exponents that depend only by the
dimensionality of the space and the magnetic order parameter.
Another fecund line of enquiry was “Does a single impurity in a metal bear a
magnetic moment?” This was related to Jun Kondo’s formulation of a problem
concerning the scattering of electrons by magnetic impurities in metals and its
eventual solution in 1980. In the presence of antiferromagnetic coupling between
an impurity and the conduction electrons of a metallic host, the combination enters
a nonmagnetic ground state below the Kondo temperature TK . The Kondo effect
is characterized by a minimum in the electrical resistivity. The study of magnetic
impurities in metals focused attention on the relation between magnetism and
electronic transport, which has proved extremely fruitful, leading to several Nobel
Prizes and the emergence in the 1990s of spin electronics.
1 History of Magnetism and Basic Concepts 35

The exchange interaction between two dilute magnetic impurities in a metal


is long-range, decaying as 1/r3 while oscillating in sign between ferrromagnetic
and antiferromagnetic, where r is their separation. The following is the Ruderman-
Kittel-Kasuya-Yosida (RKKY) exchange interaction

J (r) = aJsd 2 (sinξ − ξcosξ) /ξ4 (18)

where a is a constant, Jsd is the exchange coupling between the localized impurity
and the conduction electrons, and ξ is twice the product of r and the Fermi
wavevector. It was studied intensively in the 1970s in dilute alloys such as AuFe
or CuMn, known as spin glasses (the host is in bold type, and the impurity in
italics). The impurity in these hosts retains its moment at low temperatures, and
the RKKY exchange coupling J (∇) between a pair of spins is as likely to be
ferromagnetic (positive) as antiferromagnetic (negative). The impurity spins freeze
progressively in random orientations around a temperature Tf that is proportional to
the magnetic concentration. The nature of this transition to the frozen spin glass state
was exhaustively debated. A related issue, the long-range exchange interactions
associated with the ripples of spin polarization created by a magnetic impurity in
a metal, led to an understanding of complex magnetic order in the rare earth metals
( Chap. 14, “Magnetism of the Elements”).
The magnetism of electronic model systems such as a chain of 1s atoms with an
on-site coulomb repulsion U when two electrons occupy the same site, formulated
by John Hubbard in 1963, has proved to be remarkably complex. Control parameters
in the Hubbard model are the band filling and the ratio of U to the bandwidth, and
they lead to insulating and metallic, ferromagnetic, and antiferromagnetic solutions.

Amorphous Magnets

An important question, related to the dilute spin glass problem, was what effect does
atomic disorder have on magnetic order and the magnetic phase transition in mag-
netically concentrated systems? Here a dichotomy emerges between ferromagnetic
and antiferromagnetic interactions. The answer for materials with ferromagnetic
exchange and a weak local electrostatic (crystal field) interaction is that the atomic
disorder has little effect.
Techniques for rapidly cooling eutectic melts at rates of order 106 Ks−1
developed around 1970 produced a family of useful amorphous ferromagnetic alloys
based on Fe, Co, and Ni, with a minor amount of metalloid such as B, P, or Si. These
metallic glasses, frequently in the form of thin ribbons obtained by melt spinning,
were magnetically soft and proved that ferromagnetic order could exist without a
crystal lattice. There are no crystal axes, and weak local anisotropy due to the local
electrostatic interactions averages out. The magnetic metallic glasses are mechani-
cally strong and have found applications in transformer cores and security tags.
36 J. M. D. Coey

Amorphous materials with antiferromagnetic interactions are qualitatively dif-


ferent. Whenever the superexchange neighbors in oxides or other insulating com-
pounds form odd-membered rings, these interactions are frustrated. No collinear
magnetic configuration is able to satisfy them all. In crystalline antiferromag-
nets like rocksalt-structure NiO, the partial frustration leads to a reduced Néel
temperature, but in fully frustrated pyrochlore-structure compounds, for exam-
ple, the Néel point is completely suppressed. In the amorphous state, however,
frustration has a spatially random aspect, and it leads to random spin freezing
with a tendency to antiferromagnetic nearest-neighbor correlations, known as
speromagnetism.
The situation for amorphous rare earth intermetallic alloys, which are best
prepared by prepared by rapid sputtering, is different. There the local anisotropy
at rare earth sites is strong, and does not average out, but it tends to pin the rare
earth moments to randomly oriented easy axes in directions that are roughly parallel
to that of the local magnetization of the 3d ferromagnetic sublattice for the light rare
earths and roughly antiparallel to it for the heavy rare earths. The sign of the 3d-4f
coupling changes in the middle of the series, so that amorphous Gd-Fe alloys, for
example, are ferrimagnetic. (Gd is the case where there are no orbital moment and
no magnetocrystalline anisotropy on account of its half-filled, 4f7 shell.)
Rapid quenching can also be used to produce nanocrystalline material with
isotropic crystallite orientations of nanocrystals embedded in an amorphous matrix.
Certain soft magnetic materials have such a two-phase structure. Nanocrystalline
Nd-Fe-B produced by rapid quenching shows useful coercivity due to domain wall
pinning at the Nd2 Fe14 B nanocrystallite boundaries, but the remanence is only about
half the saturation magnetization on account of the randomly directed easy axes
of the tetragonal crystallites. The magnitude of the anisotropy and the nanoscale
dimension are critical for the averaging that determines the magnetic properties.

Magnetic Fine Particles

An early approach to the difficult problem of calculating hysteresis was to focus on


magnetization reversal in single-domain particles that were too small to benefit from
any reduction in their energy by forming a domain wall. Edmund Stoner and Peter
Wohlfarth proposed an influential model in 1948. The particles were assumed each
to have a single anisotropy axis, and the reverse field parallel to the axis necessary
for magnetic reversal was the anisotropy field Ha = 2Ku /μ0 Ms , potentially a very
large value. There was no coercivity when the field was applied perpendicular to the
axis. Insights arose from the substantial deviation of real systems from the idealized
Stoner-Wohlfarth model.
Meanwhile, the following year Néel, seeking to understand the remanent mag-
netism and hysteresis of baked clay and igneous rocks, proposed a model of
thermally driven fluctuations of the magnetization of nanometer-sized ferromagnetic
particles of volume V, a phenomenon known as superparamagnetism. The fluctua-
tion time depended exponentially on the ratio of the energy barrier to magnetic
1 History of Magnetism and Basic Concepts 37

reversal reversal  ≈ Ku V to the thermal energy kB T. Here Ku is the uniaxial


anisotropy (Eq. 16) of shape or magnetocrystalline origin. The expression for the
time τ that elapses before a magnetic reversal is

τ = τ0 exp (/kB T ) (19)

where the attempt frequency 1/τ0 was taken to be the natural resonance frequency,
∼109 Hz. When the particles are superparamagnetic, the magnetization of particles
smaller than a critical size fluctuates rapidly above a critical blocking temperature.
The magnetization at lower temperatures, or for larger particles, does not fluctuate
on the measurement timescale, and the particles are then said to be blocked. The
blocking criterion for magnetic measurements at room temperature is defined,
somewhat arbitrarily, as /kB T ≈ 25, corresponding to τ ≈ 100 s and  ≈ 1 eV
(see  Chap. 20, “Magnetic Nanoparticles”). The 10-year stability criterion is
/kB T ≈ 40. Cooling an ensemble of particles through the blocking temperature
Tb = Ku V/25kB in a magnetic field leads to a relatively stable thermoremanent
magnetization. The typical size of iron oxide particles that are superparamagnetic at
room temperature is 10 nm.
The magnetization of baked clay becomes blocked on cooling through Tb in the
Earth’s magnetic field. From the direction of the thermoremanent magnetization
of appropriately dated hearths of pottery kilns, records of the historical secular
variation of the Earth’s field could be established, a topic known as archeomag-
netism. Application of the same idea of thermoremanent magnetization to cooling
of igneous rocks in the Earth’s field provided a direct and convincing argument
for geomagnetic reversals and continental drift; rocks cooling at different periods
experienced fields of different polarities (Fig. 16), which followed an irregular
sequence on a much longer timescale than the secular variation. The reversals could
be dated using radioisotope methods on successive lava flows. This gave birth to the
subfield of paleomagnetism and in turn allowed dating of the patterns of remanent
magnetization picked up in oceanographic surveys conducted in the 1960s that
established the reality of seafloor spreading. The theory of global plate tectonics
has had far-reaching consequences for Earth science [29].
Superparamagnetic particles have found other practical uses. Ferrofluids, the
colloidal suspensions of nanoparticles in oil or water with surfactants to inhibit
agglomeration, are just one. They behave like anhysteretic ferromagnetic liquids.
Individual particles or micron-sized polymer beads loaded with many of them
may be functionalized with streptavidin and used as magnetic labels for specific
biotin-tagged biochemical species, enabling them to be detected magnetically and
separated by high-gradient magnetic separation based on the Kelvin force on a
particle with moment m, fK = (m.∇)B. Medical applications of magnetic fine
particles include hyperthermia (targeted heating by exposure to a high-frequency
magnetic field) and use as contrast agents in magnetic resonance imaging. However
the most far-reaching application of magnetic nanoparticles so far has been in
magnetic recording.
38 J. M. D. Coey

Fig. 16 Polarity of the thermoremanent magnetization measured across the floor of the Atlantic
ocean (left). Current polarity is dark; reversed polarity is light. The pattern is symmetrical about the
mid-ocean ridge, where new oceanic crust is being created. Random reversals of the Earth’s field
over the past 5 My, which are dated from other igneous lava flows, determine the chronological
pattern (right) that is used to determine the rate of continental drift, of order centimeters per year.
(McElhinney, Palaeomagnetism and Plate Tectonics [29], courtesy of Cambridge University Press)

Magnetic Recording

Particulate magnetic recording enjoyed a heyday that lasted over half a century,
beginning with analog recording on magnetic tapes in Germany in the 1930s through
digital recording on the hard and floppy discs that were introduced in the 1950s and
1960s, before eventually being superseded by thin-film recording in the late 1980
[27]. Particulate magnetic recording [30] was largely based on acicular particles of
γFe2 O3 often doped with 1–2% Co. Elongated iron particles were also used, and
acicular CrO2 was useful for rapid thermoremanent reproduction of videotapes on
account of its low Curie temperature. Magnetic digital tape recording with hard
ferrite particulate media continues to be used for archival storage.
The trend with magnetic media has always been to cram ever more digital
data onto ever smaller areas. This has been possible because magnetic recording
technology is inherently scaleable since reading is done by sensing the stray
field of a patch of magnetized particles. It follows from Eq. 2 that since the
dipole field decays as 1/r3 and the moment m ∼ Mr3 , the magnitude of B is
unchanged when everything else shrinks by the same scale factor – at least until
the superparamagnetic limit KV/kB T ≈ 40 is reached, at which point the magnetic
records become thermally unstable. To continue the scaling to bit sizes below
1 History of Magnetism and Basic Concepts 39

Fig. 17 Exponential growth of magnetic recording density over 50 years. The lower panel shows
the magnetized magnetic medium with successive generations of read heads based on anisotropic
magnetoresistance (AMR), giant magnetoresistance (GMR), and tunnel magnetoresistance (TMR)

100 nm, granular films of a highly anisotropic tetragonal Fe-Pt alloy are used
to maintain stability of the magnetic records on ever-smaller oriented crystalline
grains. The individual grains are less than 8 nm in diameter. Over the 65-year history
of hard disc magnetic recording, the bit density has increased by eight orders of
magnitude, at ever-decreasing cost (Fig. 17). Copies cost virtually nothing, and the
volume of data stored on hard discs in computers and data centers doubles every
year, so that as much new data is recorded each year as was ever recorded in all
previous years of human history. This data explosion is unprecedented, and the
third magnetic revolution, the big data revolution, is sure to have profound social
and economic consequences. Although flash memory has displaced the magnetic
hard discs from personal computers. The huge data centres, which are the physical
embodiment of the ‘cloud’ where everything we download from the interenet is
stored continue to use hard disc drives.
40 J. M. D. Coey

Methods of Investigation

Magnetism is an experimental science, and progress in understanding and applica-


tions is generally contingent on advances in fabrication and measurement technol-
ogy, whether it was fourteenth-century technology to fabricate a lodestone sphere
or twenty-first-century technology to prepare and pattern a 16-layer thin-film stack
for a magnetic sensor. The current phase of information technology relies largely on
semiconductors to process digital data and on magnets for long-term storage.
For many physical investigations, magnetic materials are needed in special
forms such as single crystals or thin films. Crystal growers have always been
assiduously cultivated by neutron scatterers and other condensed matter physicists.
Only with single crystals can tensor properties such as susceptibility, magnetostric-
tion, and magnetotransport be measured properly. Nanoscale magnetic composites
have extended the range of magnetic properties available in both hard and soft
magnets. After 1970, thin-film growth facilities (sputtering, electron beam evap-
oration, pulsed laser deposition, molecular beam epitaxy) began to appear in
magnetism laboratories worldwide. Ultra-high vacuum has facilitated the study of
surface magnetism at the atomic level, while some of the motivation to investigate
magneto-optics or magnetoresistance of metallic thin films, especially in thin-film
heterostructures, arose from the prospect of massively improved magnetic data
storage. Experimental methods are discussed in the chapters in Part 3 of this
Handbook.

Materials Preparation

Silicon steel has been produced for electromagnetic applications by hot rolling
since the beginning of the twentieth century. Annual production is now about 15
million tonnes, half of it in China. Permanent magnets, soft ferrites, and specialized
magnetic alloys are produced in annual quantities ranging from upward of a hundred
to a million tonnes. All such bulk applications of magnetism are highly sensitive to
the cost of raw materials. This effectively disqualifies about a third of the elements
in the periodic table and half of the heavy transition elements from consideration
as alloy additives in bulk material. Newer methods such as mechanical alloying
of elemental powders and rapid quenching from the melt by strip casting or melt
spinning have joined the traditional methods of high-temperature furnace synthesis
of bulk magnetic materials.
The transformation of magnetic materials science that has gathered pace since
1970 has been triggered by the ability to prepare new materials for magnetic devices
in thin-film form. The minute quantity of material needed for a magnetic sensor
or memory element, where the layers are tens of nanometers thick, means that
any useful stable element can be considered. Platinum, for example, may sell for
$30,000 per kilogram, yet it is an indispensable constituent of the magnetic medium
in the 400 million hard disc drives shipped each year that sell for about $60 each.
1 History of Magnetism and Basic Concepts 41

Uniform magnetic thin films down to atomic-scale thicknesses are produced in


many laboratories by e-beam evaporation, sputtering, pulsed laser deposition, or
molecular beam epitaxy, and the more complex tools needed to make patterned
multilayer nanometer-scale thin-film stacks are quite widely available in research
centers, as well as in the fabs of the electronics industry, which deliver the hardware
on which the technology for modern life depends.

Experimental Methods

Advances in experimental observation underpin progress in conceptual understand-


ing and technology. The discovery of magnetic resonance, the sharp absorption
of microwave or radiofrequency radiation by Zeeman split levels of the magnetic
moment of an atom or a nucleus in a magnetic field, or the collective precession
of the entire magnetic moment of a solid was a landmark in modern magnetism.
Significant mainly for the insight provided into solids and liquids at an atomic
scale, electron paramagnetic resonance (EPR) was discovered by Yevgeny Zavoisky
in 1944, and Felix Bloch and Edward Purcell established the existence of nuclear
magnetic resonance (NMR) 2 years later. In 1958, Rudolf Mössbauer discovered a
spectroscopic variant making use of low-energy gamma rays emitted by transitions
from the excited states of some stable isotopes of iron (Fe57 ) and certain rare earths
(Eu151 , Dy161 , etc.). All except Zavoisky received a Nobel Prize. The hyperfine
interactions of the multipole moments of the nuclei (electric monopole, magnetic
dipole, nuclear quadrupole) offered a point probe of electric and magnetic fields at
the heart of the atom.
Larmor precession of the total magnetization of a ferromagnet in its internal
field, usually in a resonant microwave cavity, was discussed theoretically by Landau
and Evgeny Lifshitz in 1935, and ferromagnetic resonance (FMR) was confirmed
experimentally 10 years later.
Of the non-resonant experimental probes, magnetic neutron scattering has
probably been the most influential and generally useful. A beam of thermal neutrons
from a nuclear reactor was first exploited for elastic diffraction in the USA in 1951
by Clifford Shull and Ernest Wohlan, who used the magnetic Bragg scattering
to reveal the antiferromagnetic order in MnO. Countless magnetic structures
have been determined since, using the research reactors at Chalk River, Harwell,
Brookhaven, Grenoble, and elsewhere. Magnetic excitations can be characterized by
inelastic scattering of thermal neutrons, with the help of the triple-axis spectrometer
developed in Canada by Bertram Brockhouse at Chalk River in 1956. Complete
spin-wave dispersion relations provide a wealth of information on anisotropy and
exchange. Newer accelerator-based neutron spallation sources at ISIS, Oak Ridge,
and Lund provide intense pulses of neutrons by collision of highly energetic protons
with a target of a heavy metal such as tungsten or mercury. They are most useful
for studying magnetization dynamics. The low neutron scattering and absorption
cross sections of most stable isotopes mean that neutrons can penetrate deeply into
condensed matter.
42 J. M. D. Coey

Besides neutrons, other intense beams of particles or electromagnetic radiation


available at large-scale facilities have proved invaluable for probing magnetism. The
intense, tunable ultraviolet and X-ray radiation from synchrotron sources allows
the measurement of magnetic dichroism from deep atomic levels and permits the
separate determination of spin and orbital contributions to the magnetic moment.
The spectroscopy is element-specific and distinguishes different charge states of the
same element. Spin-sensitive angular-resolved photoelectron spectroscopy makes it
possible to map the spin-resolved electronic band structure. Muon methods are more
specialized; they depend on the Larmor precession of short-lived (2.20 μs) positive
muons when they are implanted into interstitial sites in a solid. Magnetic scattering
methods are discussed in  Chap. 25, “Magnetic Scattering.” The specialized
instruments accessible at large-scale facilities supplement the traditional benchtop
measurement capabilities of research laboratories.
Perhaps the most versatile and convenient of these, used to measure the magne-
tization and susceptibility of small samples, is the vibrating sample magnetometer
invented by Simon Foner in 1956 and now a workhorse in magnetism laboratories
across the world. The sample is vibrated in a uniform magnetic field, produced by
an electromagnet or a superconducting coil, about the center of a set of quadrupole
pickup coils, which provide a signal proportional to the magnetic moment. Since
sample mass rather than sample volume is usually known, it is generally the mass
susceptibility χ m = χ /ρ that is determined.
Superconducting magnets now provide fields of up to 20 tesla or more for
NMR and general laboratory use. The 5–10 T magnets are common, and they
are usually cooled by closed-cycle cryocoolers to avoid wasting helium. Coupled
with superconducting SQUID sensors, ultrasensitive magnetometers capable of
measuring magnetic moments of 10−10 Am2 or less are widely available. (The
moment of a 5 × mm2 ferromagnetic monolayer is of order 10−8 Am2 .)
High magnetic fields, up to 35 T, require expensive special installations with
water-cooled Bitter magnets consuming many megawatts of electrical power. Resis-
tive/superconducting hybrids in Tallahassee, Grenoble and Tsukuba, and Nijmegen
can generate steady fields in excess of 40 T. Higher fields imply short pulses;
the higher the field, the shorter the pulse. Reusable coils generate pulsed fields
approaching 100 T in Los Alamos, Tokyo, Dresden, Wuhan, and Toulouse.
Magnetic domain structures are usually imaged by magneto-optic Kerr
microscopy, magnetic force microscopy, or scanning electron microscopy, although
scanning SQUID and scanning Hall probe methods have also been developed.
The Bitter method with a magnetite colloid continues to be used. All these
methods image the surface or the stray field near the surface. Ultra-fast, picosecond
magnetization dynamics are studied by optical pulse-probe methods based on the
magneto-optic Kerr effect (MOKE). Transmission electron microscopy reveals the
atomic structures of thin films and interfaces with atomic-scale resolution, while
Lorentz microscopy offers magnetic contrast and holographic methods are able to
image domains in three dimensions. Atomic-scale resolution can be achieved by
point-probe methods with magnetic force microscopy or spin-polarized scanning
tunnelling microscopy. The shift of focus in magnetism toward thin films and
1 History of Magnetism and Basic Concepts 43

thin-film devices has been matched by the development of the sensitive analytical
methods needed to characterize them. Hysteresis in thin films is conveniently
measured by MOKE or by anomalous Hall effect (AHE) when the films are
magnetized perpendicular to their plane. Magnetic fields and measurements are
discussed in  Chap. 22, “Magnetic Fields and Measurements” and other chapters
in Part 3.
An important consequence of the increasing availability of commercial super-
conducting magnets from the late 1960s was the development of medical diagnostic
imaging of tissue based on proton relaxation times measured by NMR. Thousands
of these scanners in hospitals across the world provide doctors with images of the
hearts, brains, bones, and every sort of tumor.

Computational Methods

After about 1980, computer simulation began to emerge as a third force, besides
experiment and theory, to gain insight into the physics of correlated electrons in
magnetic systems. Contributions are mainly in two areas. One is calculation of
the electronic structure, magnetic structure, magnetization, Curie temperature, and
crystal structure of metallic alloys and compounds by using the density functional
method. Magnetotransport in thin-film device structures can also be calculated. Here
there is potential to seek and evaluate new magnetic phases in silico, before trying
to make them in the laboratory. This magnetic genome program is in its infancy;
success with magnetic materials to date has been limited, but the prospects are
enticing.
The other area where computation has become a significant source of new insight
is micromagnetic simulation. The domain structure and magnetization dynamics of
magnetic thin-film structures and model heterostructures are intensely studied, both
in industrial and academic laboratories. Simulation overcomes the surface limitation
of experimental domain imaging. Software is generally based on finite element
methods or the Landau-Lifshitz-Gilbert equation for magnetization dynamics.

Spin Electronics

As technology became available in the 1960s and 1970s to prepare high-quality


metallic films with thicknesses in the nanometer range, interest in their magne-
tostansport properties grew. The terrain was being prepared for the emergence of a
new phase of research that has grown to become the dominant theme in magnetism
today – spin electronics. Spin electronics is the science of electron spin transport in
solids. Many chapters in the Handbook deal with its various aspects.
For a long time, conventional electronics treated electrons simply as elemen-
tary Fermi-Dirac particles carrying a charge e, but it ignored their spin angular
momentum ½. At first this was entirely justified; charge is conserved – the electron
has no tendency to flip between states with charge ± e, no matter how strongly
44 J. M. D. Coey

it is scattered. But angular momentum is not conserved, and spin flip scattering
is common in metals. Perhaps one scattering event in 100 changes the electron
spin state, so the spin diffusion length ls should be about ten times the mean free
path λ of the electron in a solid. When electronic device dimensions were many
microns, there was no chance of an electron retaining the memory of any initial spin
polarization it may have had, unless the device itself was ferromagnetic. Anisotropic
magnetoresistance, where the scattering depends slightly on the relative orientation
of the current and magnetization because of spin-orbit coupling, can be regarded
as the archetypical spin electronic process. The relative magnitude of effect in
permalloy, for instance, is only ∼2%, but the alloy is extremely soft, on account
of simultaneously vanishing anisotropy and magnetostriction, so a permalloy strip
with current flowing at 45◦ to the magnetic easy axis along the strip for maximum
sensitivity – which can be achieved by a superposed “barber pole” pattern of highly
conducting gold – makes a simple, miniature sensor for low magnetic fields, with
a reasonable signal-to-noise ratio. AMR sensors replaced inductive sensors in the
heads used to read data from hard discs in 1990, and the annual rate of increase of
storage density improved sharply as a result.
Meanwhile, research activity on thin-film heterostructures where the layer
thickness was comparable to the spin diffusion length began to pick up as more
sophisticated thin-film vacuum deposition tools were developed. Spin diffusion
lengths are 200 nm in Cu, or about ten times the mean free path, as expected, but
they are shorter in the ferromagnetic elements and sharply different for majority-
and minority-spin electrons. The mean free path for minority-spin electrons in Co
is only 1 nm. Particularly influential and significant was the work carried out in
1988 in the groups of Peter Grunberg in Germany and Albert Fert in France on
multilayer stacks of ferromagnetic and nonferromagnetic elements that led to the
discovery of giant magnetoresistance (GMR). The effect depended on electrons
retaining some of their spin polarization as they emerged from a ferromagnetic
layer and crossed a nonmagnetic layer before reaching another ferromagnetic layer.
Big changes of resistance were found when the relative alignment of the adjacent
ferromagnetic iron layers in an Fe-Cr multilayer stack was altered from antiparallel
to parallel by applying a magnetic field (Fig. 18). At first, large magnetic fields and
low temperatures were needed to see the resistance changes, but the structure was
soon simplified to a sandwich of just two ferromagnetic layers with a copper spacer
that became known as a spin valve. Spin valves worked at room temperature, and
they were sensitive to the small stray fields produced by recorded magnetic tape or
disc media. In order to make a useful sensor, it was necessary to pin the direction
of magnetization of one of the ferromagnetic layers while leaving the other free to
respond to an in-plane field (Fig. 19).
It was here that the phenomenon of exchange bias came to the rescue. First
discovered in Co/CoO core shell particles by Meiklejohn and Bean in 1956, it was
extended to antiferromagnetic/ferromagnetic thin-film pairs in Néel’s laboratory
in Grenoble in the 1960s. By pinning one ferromagnetic layer with an adjacent
antiferromagnet (initially NiO), a useful GMR sensor could be produced with
a magnetoresistance change of order 10%. Exchange-biased GMR read heads
1 History of Magnetism and Basic Concepts 45

Fig. 18 Original measurement of giant magnetoresistance of a FeCr multilayer stack, where


the iron layers naturally adopt an antiparallel conduction, which can be converted to a parallel
configuration in an applied field [31]

developed by Stuart Parkin and colleagues went into production at IBM in 1998 – a
remarkably rapid transfer from a laboratory discovery to mass production. Exchange
bias was the first practical use of an antiferromagnet. The Nobel Physics Prize was
awarded to Fert and Grunberg for their work in 2007.
Subsequent developments succeeded in eliminating the influence of the stray
field of the pinned layer on the free layer by means of a synthetic antiferromagnet.
This was another sandwich stack, like the slimmed-down spin valve, except the
spacer was not copper, but an element that transferred exchange coupling from one
ferromagnetic layer to the other. Ruthenium proved to be ideal, and a layer just
0.7 nm thick was found to be ideal for antiferromagnetic coupling [32].
GMR’s tenure as read-head technology was to prove as short-lived as that of
AMR. A new pretender with a much larger resistance change was based on the
magnetic tunnel junction (MTJ), a modified spin valve where the nonmagnetic
metal spacer is replaced by a thin layer of nonmagnetic insulator. Electron tunneling
across an atomically thin vacuum barrier had been a striking prediction of quantum
mechanics implicit in the idea of the wavefunction. The thin barrier was at first made
of amorphous alumina, but it was replaced by crystalline MgO after it was found in
2004 that junctions where the MgO barrier acts as a spin filter exhibit tunneling
magnetoresistance (TMR) in excess of 200% [33, 34] (Fig. 19). The adoption of
46 J. M. D. Coey

B B
I
free free
pinned pinned
af af

ΔR
I
ΔR
MR = (R↑↓−R↑↑)/R↑↑

B B
Spin valve sensor Magnetic tunnel junction (MTJ)
Fig. 19 Magnetic bilayer spin-valve stacks used as sensor (left) or as a memory element (right). In
each case, the magnetization lies in-plane, and the lower ferromagnetic reference layer is pinned by
exchange bias with the purple underlying antiferromagnetic layer, while the upper ferromagnetic
free layer changes its orientation in response to the applied magnetic field. The change in stack
resistance is plotted as a function of applied field. The magnetoresistance ratio MR is defined as the
normalized resistance change between parallel and antiparallel orientation of the two ferromagnetic
layers

TMR sensors in read heads in 2005 was accompanied by a change from in-plane to
perpendicular recording on the magnetic medium.
Despite the changing generations of readers, the hard disc writer remained
what is always had been, a miniature electromagnet that delivers sufficient flux
to a patch of magnetic medium to overcome its coercivity and write the record.
The extreme demands of magnetic recording have driven contactless magnetic
sensing to new heights of sensitivity and miniaturization requiring increasingly
hard magnetic media and new ways of writing them. Thin-film GMR and TMR
structures have also taken a new life as magnetic switches for nonvolatile memory
and logic. Most prominent is magnetic random access memory (MRAM), where
huge arrays of memory cells are based on magnetic tunnel junctions. Magnetic
sensing is discussed in  Chaps. 31, “Magnetic Sensors,” and  22, “Magnetic
Fields and Measurements.”
Magnetic thin-film technology has now advanced to the point where uniform
layers in synthetic antiferromagnets and magnetic tunnel junctions only a few atoms
thick are routinely deposited on entire 200 or 300 mm silicon wafers. A corollary
of the short spin diffusion length of electrons in metals is the short distance – a
few atomic monolayers – necessary for an electron to acquire spin polarization on
transiting a ferromagnetic layer. Spin-polarized electron currents are central to spin
electronics.
1 History of Magnetism and Basic Concepts 47

The relation between magnetism and the angular momentum of electrons was
unveiled in Larmor precession and the Einstein-de Haas experiment over a hundred
years ago, but only in the present century has it become commonplace to associate
electric currents with short-range flows of angular momentum. A spin-polarized
current carrying its angular momentum into a ferromagnetic thin-film element can
exert torque in two ways. It can create an effective magnetic field, causing Larmor
precession of the magnetization of the element, and it can exert spin transfer torque,
described by John Slonczewski in 1996 that counteracts damping of the precession
and can be used to stabilize high-frequency oscillations or switch the magnetization
without the need for an external magnetic field. Spin torque switching is effective
for elements smaller than 100 nm in size, and unlike switching by current-induced
“Oersted” fields, it is scalable – an essential requirement for electronic devices. Luc
Berger showed that spin torque can also be used to manipulate domain walls.
A recurrent theme in the recent development of magnetism is the role of the spin-
orbit interaction. It is critically important in thin films [35], being responsible not
only for the Kerr effect, magnetocrystalline anisotropy, and anisotropic magnetore-
sistance but also for the anomalous Hall effect and the spin Hall effect, whereby
spin-orbit scattering of a current passing through a heavy metal or semiconductor
produces a buildup of electrons with opposite spin on opposite sides of the
conductor. This transverse spin current created by spin-orbit scattering enables the
injection of angular momentum into an adjacent ferromagnetic layer and the change
of its magnetization direction, an effect known as spin-orbit torque. Conversely,
the inverse spin Hall effect is the appearance of a voltage across the heavy metal on
pumping spin-polarized electrons into it from an adjacent ferromagnet, for example,
by exciting ferromagnetic resonance.
The origin of the intrinsic anomalous Hall effect was an open question in
magnetism, for well over a hundred years. A consensus is now building that it is due
to the geometric Berry phase acquired by electrons moving adiabatically through a
magnetic medium. The phase can be acquired from a non-collinear spin structure
in real space or from topological singularities in the band sturcture in reciprocal
space. Circular micromagnetic defects, known as skyrmions are also topologically
protected.
Another manifestation of spin-orbit interaction is the Rashba effect; when an
electric current is confined at an interface or surface, it tends to create a spin
polarization normal to the direction of current flow. One of the most remarkable
surface phenomena, arising from work by Haldane in 1988, is the possibility of
topologically protected spin currents. A special feature of the band structure ensures
that electrons at the surface or edges of some insulators or semiconductors are in
gapless states. Electrons in these states can propagate around the surface without
scattering, and they exhibit a spin order that winds around the surface as the
direction of electron spin is usually locked at right angles to their linear momentum.
Electrons at surfaces and interfaces can behave quite differently from electrons in
the bulk, and interfaces are at the heart of electronic devices. The introduction
of topological concepts into the discussion of spin-polarized electronic transport
48 J. M. D. Coey

and magnetic defects is providing new insight into magnetism at the atomic and
mesoscopic scales.

Conclusion

Magnetism since 1945 has been an area rich in discovery and useful applications,
not least because of the tremendous increase in numbers of scientists and engineers
working in the field. Magnet ownership for citizens of the developed world has
skyrocketed from 1 or 2 magnets in 1945 to 100–200 60 years later or something
of order a trillion if we count the individual magnetic bits on a hard disc in
a desktop computer. Countless citizens throughout the world during this period
already experienced magnetism’s bounty at first hand in the form of a cassette tape
recorder, and nowadays they can access the vast stores of magnetically recorded
information in huge data centers via the Internet using a handheld device.
Magnetism is therefore playing a crucial role in the big data revolution that is
engulfing us, by enabling the permanent data storage, from which we can make
instant copies at practically no cost. It may deliver more nonvolatile computer
memory if MRAM proves to a winning technology and possibly facilitate data
transfer at rates up to the terahertz regime with the help of spin torque oscillators.
There are potential magnetic solutions to the problems of ballooning energy
consumption and the data rate bottleneck. There is potential to implement new
paradigms for computation magnetically. While there is no certainty regarding the
future form of information technology, improved existing solutions often have an
inside track. Magnetism and magnetic materials may be a good bet.
There have been half a dozen paradigm shifts – radical changes in the ways of
seeing and understanding the magnet and its magnetic field – during its 2000-year
encounter with human curiosity. Implications of the big data revolution for human
society are only beginning to come into focus, but they are likely to be as profound as
on the previous two occasions when magnetism changed the world. This Handbook
is a guide to what is going on.

Acknowledgments The author is grateful to Science Foundation Ireland for continued support,
including contracts 10/IN.1/I3006, 13/ERC/I2561 and 16/IA/4534.

Appendix: Units

By the middle of the nineteenth century, it was becoming urgent to devise a standard
set of units for electrical and magnetic quantities in order to exchange precise
quantitative information. The burgeoning telegraph industry, for example, needed
a standard of electrical resistance to control the quality of electrical cables. Separate
electrostatic and electromagnetic unit systems based on the centimeter, the gram and
the second had sprung into existence, and Maxwell and Jenkin proposed combining
them in a coherent set of units in 1863. Their Gaussian cgs system was adopted
1 History of Magnetism and Basic Concepts 49

internationally in 1881. Written in this unit system, Maxwell’s equations relating


electric and magnetic fields contain explicit factors of c, the velocity of light.
Maxwell also introduced the idea of dimensional analysis in terms of the three basic
quantities of mass, length, and time. The magnetic field H and the induction B are
measured, respectively, in the numerically identical but dimensionally different units
of oersted (Oe) and gauss (G).
Another basic unit, this time of electric current, was adopted in the Système
International d’Unités (SI) in 1948. The number of basic units and dimensions in any
system is an arbitrary choice; the SI (International System of Units) uses four insofar
as we are concerned, the meter, kilogram, second, and ampere (or five if we include
the mole). The system has been adopted worldwide for the teaching of science and
engineering at school and universities; it embodies the familiar electrical units of
volt, ampere, and ohm for electrical potential, current, and resistance. Maxwell’s
equations written in terms of two electric and two magnetic fields contain no factors
of c or 4π in this system (Eq. 7), but they inevitably crop up elsewhere. B and H are
obviously different quantities. The magnetic field strength H, like the magnetization
M, has units of Am−1 . The magnetic induction B is measured in tesla (1 T ≡
1 kgs2 A−2 ). Magnetic moments have units of Am2 , clearly indicating the origin
of magnetism in electric currents and the absence of magnetic poles as real physical
entities. The velocity of light is defined to be exactly 299,792,458 ms−1 . The two
constants μ0 and ε0 , the permeability and permittivity of free space, are related by
μ0 ε0 = c2 , where μ0 was 4π 10−7 kgs−2 A−2 according to the original definition of
the ampere. However, in the new version of SI, which avoids the need for a physical
standard kilogram, the equality of μ0 and 4π 10−7 is not absolute, but it is valid to
ten significant figures.
Only two of the three fields B, H, and M are independent (Fig. 4). The relation
between them is Eq. 8, B = μ0 (H + M). This is the Sommerfeld convention for SI.
The alternative Kenelly convention, often favored by electrical engineers, defines
magnetic polarization as J = μ0 M, so that the relation becomes B = μ0 H + J. We

Table 1 Numerical conversion factors between SI and cgs units


Physical quantity Symbol SI to cgs conversion cgs to SI conversion
B-field (magnetic flux B 1 tesla = 10 kilogauss 1 gauss* = 0.1
density) millitesla
H-field (magnetic field H 1 kAm−1 = 12.57 oersted 1 oersted§ = 79.58
intensity) Am−1
Magnetic moment m 1 Am2 = 1000 emu 1 emu = 1 mAm2
Magnetization M 1 Am−1 = 12.57 gauss† 1 gauss† = 79.58 Am−1
Specific magnetization σ 1 Am2 kg−1 = 1 emu g−1 1 emu g−1 = 1
Am2 kg−1
Magnetic energy density (BH) 1 kJm−3 = 0.1257 MGOe 1 MGOe = 7.96 kJm−3
Dimensionless χ 1 (SI) = 1/4π (cgs) 1 (cgs) = 4π (SI)
susceptibility M/H
*symbol G; § symbol Oe; † 4πM; Note: 12.57 = 4π; 79.58 = 1000/4π
50 J. M. D. Coey

follow the Sommerfeld convention in this Handbook. The magnetic field strength H
is not measured in units of Tesla in any generally accepted convention, but it can be
so expressed by multiplying by μ0 .
At the present time, Gaussian cgs units remain in widespread use in research
publications, despite the obvious advantages of SI. The use of the cgs system in
magnetism runs into the difficulty that units of B and H, G and Oe, are dimensionally
different but numerically the same; μ0 = 1, but it normally gets left out of the
equations, which makes it impossible to check whether the dimensions balance.
Table 1 lists the conversion factors and units in the two systems. The cgs equivalent
of Eq. 8 is B = H + 4πM. The cgs unit of charge is defined in such a way that
ε0 = 1/4πc and μ0 = 4π/c so factors of c appear in Maxwell’s equations in place
of the electric and magnetic constants. Convenient numerical conversion factors
between the two systems of units are provided in Table 1.
Theoretical work in magnetism is sometimes presented in a set of units where
c =  = kB = 1. This simplifies the equations, but does nothing to facilitate
quantitative comparison with experimental measurements.

References
1. Kloss, A.: Geschichte des Magnetismus. VDE-Verlag, Berlin (1994)
2. Matthis, D.C.: Theory of Magnetism, ch. 1. Harper and Row, New York (1965)
3. Needham, J.: Science and Civilization in China, vol. 4, part 1. Cambridge University Press,
Cambridge (1962)
4. Schmid, P.A.: Two early Arabic sources of the magnetic compass. J. Arabic Islamic Studies. 1,
81–132 (1997)
5. Fowler, T.: Bacon’s Novum Organum. Clarendon Press, Oxford (1878)
6. Pierre Pèlerin de Maricourt: The Letter of Petrus Peregrinus on the Magnet, AD 1292, Trans.
Br. Arnold. McGraw-Hill, New York (1904)
7. Gilbert, W: De Magnete, Trans. P F Mottelay. Dover Publications, New York (1958)
8. Fara, P.: Sympathetic Attractions: Magnetic Practices, Beliefs and Symbolism in Eighteenth-
Century England. Princeton University Press, Princeton (1996)
9. Mottelay, P.F.: Bibliographical History of Electricity and Magnetism. Arno Press, New York
(1975)
10. Faraday, M.: Experimental Researches in Electricity, volume III. Bernard Quartrich, London
(1855)
11. Maxwell, J.C.: A Treatise on Electricity and Magnetism, two volumes. Clarendon Press,
Oxford (1873) (Reprinted Cambridge University Press, 2010)
12. Bertotti, G.: Hysteresis in Magnetism. Academic Press, New York (1998)
13. Brown, W.F.: Micromagnetics. Interscience, New York (1963)
14. Sato, M., Ishii, Y.: Simple and approximate expressions of demagnetizing factors of uniformly
magnetized rectangular rod and cylinder. J. Appl. Phys. 66, 983–988 (1989)
15. Hunt, B.J.: The Maxwellians. Cornell University Press, New York (1994)
16. Joule, J.P.: On the Effects of Magnetism upon the Dimensions of Iron and Steel Bars.
Philosoph. Mag. Third Series. 76–87, 225–241 (1847)
17. Thomson, W.: On the electrodynamic qualities of metals. Effects of magnetization on the
electric conductivity of nickel and iron. Proc. Roy. Soc. 8, 546–550 (1856)
18. Kerr, J.: On rotation of the plane of the polarization by reflection from the pole of a magnet.
Philosoph. Mag. 3, 321 (1877)
1 History of Magnetism and Basic Concepts 51

19. Ewing, J.A.: Magnetic Induction in Iron and Other Metals, 3rd edn. The Electrician Publishing
Company, London (1900)
20. Tomonaga, S.: The Story of Spin. University of Chicago Press, Chicago (1974)
21. Marage, P., Wallenborn, G. (eds.): Les Conseils Solvay et Les Débuts de la Physique Moderne.
Université Libre de Bruxelles (1995)
22. Ballhausen, C.J.: Introduction to Ligand Field Theory. McGraw Hill, New York (1962)
23. Bozorth, R.M.: Ferromagnetism. McGraw Hill, New York (1950) (reprinted Wiley – IEEE
Press, 1993)
24. Goodenough, J.B.: Magnetism and the Chemical Bond. Interscience, New York (1963)
25. Smit, J., Wijn, H.P.J.: Ferrites; Physical Properties of Ferrrimagnetic Oxides. Philips Technical
Library, Eindhoven (1959)
26. Coey, J.M.D., Viret, M., von Molnar, S.: Mixed valence manganites. Adv. Phys. 48, 167 (1999)
27. Wang, S.X., Taratorin, A.M.: Magnetic Information Storage Technology. Academic Press, San
Diego (1999)
28. Coey, J.M.D. (ed.): Rare-Earth Iron Permanent Magnets. Clarendon Press, Oxford (1996)
29. McElhinney, M.W.: Palaeomagnetism and Plate Tectonics. Cambridge University Press (1973)
30. Daniel, E.D., Mee, C.D., Clark, M.H. (eds.): Magnetic Recording, the First Hundred Years.
IEEE Press, New York (1999)
31. Baibich, M.N., Broto, J.M., Fert, A., Nguyen Van Dau, F., et al.: Giant magnetoresistance of
(001)Fe/(001)Cr magnetic superlattices. Phys. Rev. Lettters. 61, 2472 (1988)
32. Parkin, S.S.P.: Systematic variation on the strength and oscillation period of indirect magnetic
exchange coupling through the 3d, 4d and 5d transition metals. Phys. Rev. B. 67, 3598 (1991)
33. Parkin, S.S.P., Kaiser, C., Panchula, A., Rice, P.M., Hughes, B., et al.: Giant tunneling
magnetoresistance with MgO (100) tunnel barriers. Nat. Mater. 3, 862–867 (2004)
34. Yuasa, S., Nagahama, T., Fukushima, A., Suzuki, Y., Ando, K.: Giant room-temperature
magnetoresistance in single-crystal Fe/MgO/Fe magnetic tunnel junctions. Nat. Mater. 3,
868–871 (2004)
35. Sinova, J., Valenzuela, S.O., Wunderlich, J., Bach, C.H., Wunderlich, J.: Spin hall effects. Rev.
Mod. Phys. 87, 1213 (2015)

Michael Coey received his PhD from the University of Manitoba


in 1971; he has worked at the CNRS, Grenoble, IBM, York-
town Heights, and, since 1979, Trinity College Dublin. Author
of several books and many papers, his interests include amor-
phous and disordered magnetic materials, permanent magnetism,
oxides and minerals, d0 magnetism, spin electronics, magneto-
electrochemistry, magnetofluidics, and the history of ideas.
Magnetic Exchange Interactions
2
Ralph Skomski

Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
Quantum-Mechanical Origin of Exchange . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
One-Electron Wave Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
Electron-Electron Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
Stoner Limit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
Correlations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
Heisenberg Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
Hubbard Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
Specific Exchange Mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
Intra-Atomic Exchange . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
Indirect Exchange . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
Itinerant Exchange . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
Bethe-Slater Curve . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
Metallic Correlations and Kondo Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
Exchange and Spin Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
Curie Temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
Magnetic Order and Noncollinearity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
Spin Waves and Anisotropic Exchange . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
Experimental Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
Antiferromagnetic Spin Chains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
Dimensionality Dependence of Quantum Antiferromagnetism . . . . . . . . . . . . . . . . . . . . . . . 94
Frustration, Spin Liquids, and Spin Ice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99

R. Skomski ()
University of Nebraska, Lincoln, NE, USA
e-mail: rskomski@neb.rr.com

© Springer Nature Switzerland AG 2021 53


J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_2
54 R. Skomski

Abstract

The electrostatic repulsion between electrons, combined with quantum mechan-


ics and the Pauli principle, yields the atomic-scale exchange interaction. Intra-
atomic exchange determines the size of the atomic magnetic moments. Inter-
atomic exchange ensures long-range magnetic order and determines the ordering
(Curie or Néel) temperature. It also yields spin waves and the exchange stiffness
responsible for the finite extension of magnetic domains and domain walls.
Intra-atomic exchange determines the size of the atomic magnetic moments.
Positive and negative exchange constants mean parallel (ferromagnetic) and
antiparallel (antiferromagnetic) spin alignments. As a rule, direct exchange
and Coulomb interaction favor ferromagnetic spin structures, whereas inter-
atomic hopping tends to be ferromagnetic and is often the main consideration.
The basic interatomic exchange mechanisms include superexchange, double
exchange, Ruderman-Kittel exchange, and itinerant exchange. Exchange inter-
actions may also be classified according to specific models or phenomena.
Examples are Heisenberg exchange, Stoner exchange, Hubbard interactions,
anisotropic exchange, Dzyaloshinski-Moriya exchange, and antiferromagnetic
spin fluctuations responsible for high-temperature superconductivity. From the
viewpoint of fundamental physics, exchange interactions differ by the role of
electron correlations, the strongly correlated Heisenberg exchange and weakly
correlated itinerant exchange at the opposite ends of the spectrum. Correlations
are also important for the understanding of some exotic exchange phenomena,
such as frustration and quantum spin liquid behavior.

Introduction

Solid-state magnetism is caused by interacting atomic moments or “spins” (Fig. 1).


In the absence of such interactions, the spins would point in random directions, and
the net magnetization would be zero. Ferromagnetic (FM) order requires positive
interactions (a), which favor parallel spin alignment, ↑↑, and yields a nonzero
net magnetization. Antiferromagnetic (AFM) order (b) is caused by negative
interactions and corresponds to antiparallel spin alignment ↑↓ between neighboring
atoms. For reasons discussed below, these interactions are referred to as exchange
interactions. Aside from the interatomic exchange illustrated in Fig. 1, there are
intra-atomic exchange interactions. For example, Fe2+ ions in oxides have six 3d
electrons and the spin structure ↑↑↑↑↑↓, which yields a net atomic moment of
2 μB . Magnetic moments in transition-metal elements and alloys tend to be non-
integer, as exemplified by Ni, which has a moment of 0.61 μB per atom. Such
non-integer moments reflect the itinerant Stoner exchange, which contains both
inter- and intra-atomic contributions.
By about 1920, it had become clear that magnetostatic interactions cannot
explain ferromagnetism at and above room temperature. Weiss’ mean-field the-
ory assumes that a molecular field stabilizes ferromagnetic order. However, the
2 Magnetic Exchange Interactions 55

Fig. 1 Interatomic exchange and magnetic order: (a) antiferromagnetism (AFM) and (b) ferro-
magnetism (FM)

molecular fields required in the theory (several 100 teslas) are much higher than
typical magnetostatic interaction fields, which are only of the order of 1 tesla.
Equating the thermal energy kB T with the Zeeman energy μo μB H yields the
conversion μB /kB = 0.672 K/T, meaning that low-temperature thermal excitations
of about 1 kelvin destroy any magnetic order caused by magnetic fields of about
1 T. The weakness of Zeeman and other magnetic interactions reflects the relativistic
character of magnetism: The ratio of magnetic and electrostatic interactions is of the
order of 1/α 2 , where α = 1/137 is Sommerfeld’s fine structure constant.
Aware of the smallness of purely magnetic interactions, Werner Heisenberg
concluded in 1928 that ferromagnetic order must be of electrostatic origin, realized
on a quantum-mechanical level [1]. He found that the Coulomb repulsion between
electrons

1 e2
U (r 1 , r 2 ) = (1)
4πεo | r 1 − r 2 |

in combination with the Pauli principle yields a strong effective field consistent
with experiment. The Pauli principle forbids the occupancy of an orbital by two
electrons of parallel spin. In real space, it yields an exchange hole, that is, electrons
with parallel spins (↑↑) stay away from each other, while electrons with antiparallel
spin (↑↓) can come arbitrarily close, which carries a Coulomb-energy penalty. In a
nutshell, this is the origin of ferromagnetic exchange. A different consideration is
that even electrons of antiparallel spin (↑↓) avoid each other to some extent due to
their Coulomb repulsion, which is known as the correlation hole. The correlation
hole weakens the trend toward ferromagnetism.
Consider two electrons 1 and 2 in two atomic orbitals L (for left) and R (for
right). The real-space part of the wave function is
56 R. Skomski

1
± (r 1 , r 2 ) = √ (φL (r 1 )φR (r2 ) ± φR (r 1 )φL (r 2 )) (2)
2

where the upper and lower signs correspond to ↑↓ and ↑↑, respectively.

Using Eq. (2) to evaluate the Coulomb interaction EC = Ψ ± ∗ U Ψ ± dr1 dr2
yields an energy splitting of ±JD . The integral

±JD = φL∗ (r 1 )φR∗ (r 2 )U (r 1 , r 2 )φR (r 1 )φL (r 2 )dV1 dV2 (3)

is referred to as the exchange integral or direct exchange. If JD is positive, then


the energy of the FM state is lower than that of the AFM state, favoring ↑↑
alignment. Since direct exchange JD is of electrostatic origin, it has the right order
of magnitude to explain ferromagnetism.
However, equating the net exchange J with JD has a number of flaws. First,
JD is the electrostatic self-interaction energy of a fictitious charge distribution
ρ F (r) = – e φ ∗ L (r) φ R (r) and therefore always positive. This is at odds with exper-
iment, because antiferromagnetism is well established in many materials. Second,
Eq. (2) means that the left (L) and right (R) atoms harbor exactly one electron each
[1, 2]. This approximation, known as the Heitler-London approximation in chem-
istry, amounts to ignoring the ionic configurations φ L (r1 ) φ L (r2 ) and φ R (r1 ) φ R (r2 ).
In fact, some iconicity is expected on physical grounds, because electrons hop
between atoms and therefore temporarily create ionic configurations.
 Third, the
wave functions φ L (r) and φ R (r) exhibit some overlap So = φ L (r) φ R (r) dr. This
overlap is responsible for interatomic hopping, affecting the one-electron levels and
reducing the net exchange. Furthermore, overlap corrections diverge with increasing
number N of electrons involved, which is known as non-orthogonality catastrophe.
Sections “One-Electron Wave Functions” and “Electron-Electron Interactions”
solve the overlap problem by using Wannier-type orthogonalized orbitals.
It is nontrivial to predict magnetism from the chemical composition. For
example, MnBi, ZrZn2 , and CrBr3 are all ferromagnetic but do not contain any
ferromagnetic element. Section “Specific Exchange Mechanisms” describes a num-
ber of important exchange mechanisms in metals and insulators. It is important to
distinguish between intra-atomic exchange, which is responsible for the formation
of atomic magnetic moments, and interatomic exchange, which determines the type
of magnetic order and the ordering temperature. Examples of magnetic order are the
FM and AFM structures of Fig. 1, but there also exist noncollinear spin structures,
caused, for example, by competing exchange or Dzyaloshinski-Moriya interactions
(Sections “Curie Temperature” and “Magnetic Order and Noncollinearity”).
Beyond determining magnetic order, exchange is important in micromagnetism,
where the exchange stiffness affects the sizes of magnetic domains and domain
walls (Section “Spin Waves and Anisotropic Exchange”). Exchange is also involved
in various “exotic” magnetic systems, such as frustrated spin structures and
quantum-spin liquids (Sections “Curie Temperature” and “Magnetic Order and
Noncollinearity”).
2 Magnetic Exchange Interactions 57

Quantum-Mechanical Origin of Exchange

Exchange reflects the interplay between independent-electron level splittings (T ),


Coulomb repulsion (U ), and the direct exchange integral (JD ). This section
elaborates the fundamentals of this relationship, starting from one-electron wave
functions (Section “One-Electron Wave Functions”), introducing wave functions of
interacting electrons (Section “Electron-Electron Interactions”), and discussing a
number of fundamental limits and models (Sections “Stoner Limit”, “Correlations”,
“Heisenberg Model”, and “Hubbard Model”).

One-Electron Wave Functions

Well-separated atoms are described by atomic wave functions of on-site energy Eat ,
but in molecules and solids, the wave functions of neighboring atoms overlap and
yield interatomic hybridization. This section focuses on two s electrons in atomic
dimers, such as H2 and hypothetical Li2 (Fig. 2). Most features of this model can be
generalized to solids, although solids exhibit additional many-electron effects. The
one-electron Hamiltonian corresponding to Fig. 2 is

2 2
H.1 (r) = ∇ + VL (r) + VR (r) (4)
2me

where L and R stand for right and left, respectively, and VL/R (r) = Vo (| r – RL/R | )
are the atomic potentials. In the case of hydrogen-like atoms of radius Rat = ao /Z,
the atomic ground state has the eigenfunction φ(r)∼ exp (–r/Rat ) and the energy
Eat = Z2 e2 /8πεo ao . The vicinity of the second atom means that the atomic wave
functions φ L (r) = |Lo > and φ R (r) = |Ro > overlap, which is described by the overlap

Fig. 2 Symmetric and antisymmetric wave functions: (a) atomic wave function, (b) Wannier
function, (c) antibonding state |σ *>, and (d) bonding state |σ >
58 R. Skomski

integral

So =< Lo VR Ro > (5)

The overlap causes interatomic hopping and yields a level splitting into bonding and
antibonding states. The respective energies are Eo ± T , where T is the hopping
integral. The hybridized eigenfunctions (Fig. 2(c-d)) are |σ > ∼ |Lo > + |Ro >
(bonding) and |σ ∗ > ∼ |Lo > – |Ro > (antibonding), both having So -dependent
normalizations. The label σ refers to the ss σ -bond between the two s orbitals and
leads to T < 0 in this specific model. The on-site energy Eo differs from the atomic
energy Eat by the crystal-field energy ECF ≈ <Lo |VR |Lo >.

Traditional and Modern Analyses The determination and interpretation of the


hopping integral T require care, because hopping affects the net exchange J and
may change its sign. T is often approximated by

To = <Lo |VR |Ro > = φL∗ (r) VR (r) φR (r) dr (6)

but this interpretation is qualitative only, because the overlap correction to T is of


the same order of magnitude as To itself. Equation (6) therefore conflates the related
phenomena of hopping and wave-function overlap. This distinction is related to the
abovementioned non-orthogonality catastrophe.
Orthogonality problems are avoided by the use of orthogonalized atomic wave
functions or Wannier functions [3]. These functions are similar to atomic wave func-
tions but contain some admixture of neighboring orbitals to ensure orthogonality. In
the present model [4]

1   1  
|L> = √ |σ > + |σ ∗ > and |R> = √ |σ > − |σ ∗ > (7)
2 2

Figure 2(b) shows one of these Wannier functions. The two wave functions (c-d)
correspond to rudimentary wave vectors k = 0 (bonding) and k = π/a (antibonding).
Solids are very similar in this regard, except that k varies continuously (band
structure).
It is instructive to discuss the parameters involved for large interatomic dis-
tances R [2]. In this extreme-tight-binding limit, So = ½ (R/Rat )2 exp(–R/Rat ),
ECF = 2(Rat /R)Eat , and T = So ECF . Since ECF decreases only slowly, scaling
as 1/R, the asymptotic behavior of T is governed by the exponential decay of So .
In terms of the Wannier functions |L > and |R>, the one-electron Hamiltonian of
Eq. (4 ) assumes the very simple matrix structure
 
Eo T
H1 = (8)
T Eo
2 Magnetic Exchange Interactions 59

The diagonalization of this Hamiltonian is trivial, reproducing E± = Eo ± T


and yielding

1 1
|σ > = √ (|L> + |R>) and |σ ∗ > = √ (|L> − |R>) (9)
2 2

Equations (8, 9) remove the overlap integral from explicit consideration and
constitute a great scientific and practical simplification.

Electron-Electron Interactions

Ferromagnetism is caused by electron-electron interactions. Addition of the


= U (r , r ) to Eq. (4) yields
Coulomb energy H12 1 2

H (r 1 , r 2 ) = H1 (r 1 ) + H1 (r 2 ) + U (r 1 , r 2 ) (10)

To diagonalize this Hamiltonian, it is convenient to use two-electron wave func-


tions Ψ i constructed from Wannier functions, namely, Ψ 1 = |LL>, Ψ 2 = |LR>,
Ψ 3 = |RL>, and Ψ 4 = |RR>. Since <L|R > = 0, these functions are all orthogonal,
and Eq. (10) becomes
⎛ ⎞
U T T JD
⎜T 0 JD T ⎟
H= ⎜
⎝T
⎟ (11)
JD 0 T ⎠
JD T T U

where a physically unimportant zero-point energy has been ignored. The Coulomb
parameter U is the extra energy required to put a second electron onto a given atom
(R or L), essentially

e2 n(r 1 ) n(r 2 )
U= dr 1 dr 2 (12)
4πεo | r1 − r2 |

where n(r) = nL/R (r). Unlike JD , which decrease exponentially with interatomic
distance, U is an atomic parameter and more or less independent of crystal structure.
Both U and T tend to be large, several eV, whereas JD is rather small, typically of
the order of 0.1 eV. This indicates that JD is not the only or even the most important
contribution to interatomic exchange. For example, the exchange in the H2 molecule
is antiferromagnetic, in spite of JD being positive.
Equation (11) can be diagonalized analytically. There are two low-lying states

1 1
|↑↑> = √ |LR> − √ |RL> (13)
2 2
60 R. Skomski

cos χ sin χ
|↑↓> = √ (|LR> + |RL>) + √ (|LL> + |RR>) (14)
2 2

where tan (2χ ) = –4T /U [4]. Equation (14) is a superposition of two Slater
determinants, described by the mixing angle χ . The corresponding energy levels
are

E↑↑ = –JD (15)


U U2
E↑↓ = +D − 4T 2 + (16)
2 4
 
Defining an effective exchange as J = E↑↑ –E↑↑ /2 yields

U U2
J = JD + − T2+ (17)
4 16

This equation shows that interatomic hopping (T ) reduces the net exchange
interaction. The effect comes from the admixture of |LL> and |RR> to |LR> + |RL>
(Eq. (14)) which is ignored in Eq. (2).

Stoner Limit

In the metallic limit of strong interatomic hopping (T


U ), Eq. (17) becomes

U
J = + JD − |T | (18)
4

This equation predicts ferromagnetism for sufficiently small hopping T and roughly
corresponds to the Stoner theory [5] of itinerant transition-metal magnets (Sec-
tion “Itinerant Exchange”). Since U
JD , the driving force behind Stoner
ferromagnetism is the Coulomb integral U , not the direct exchange JD [6]. The
interatomic hopping competes against the electron-electron interactions described
by the Stoner parameter I = U /4 + JD , whereas a refined calculation for transition
metals yields I = U /5 + 1.2 Jat [7]. Here Jat is the intra-atomic exchange, which
merges with the interatomic exchange in the itinerant limit.
Equation (18) yields a very simple and scientifically successful explanation,
namely, that ferromagnetism occurs when the one-electron level splitting, ±|T | in
the model of Section “Electron-Electron Interactions”, is sufficiently small com-
pared to the nearly crystal-independent Coulomb parameter U . Figure 3 illustrates
the physics behind this mechanism. The Coulomb repulsion U favors the FM config-
uration, but the FM alignment carries a one-electron energy penalty. More precisely,
2 Magnetic Exchange Interactions 61

Fig. 3 Origin of magnetism in the independent-electron picture. The one-electron level splitting
into bonding (σ ) and antibonding (σ *) states favors ↑↓ spin pairs, whereas the Coulomb repulsion
between the two |σ > electrons yields ↑↑ coupling so long as the Coulomb energy is larger
than the one-electron level splitting. The independent-electron nature of this picture is seen from
two features. First, the electrons occupy one-electron levels (σ and σ *). Second, the Coulomb
interaction can be interpreted as an effective field (Stoner exchange field)

the “one-electron” contributions of this section are actually independent-electron


contributions treated on a quantum-mechanical mean-field level [8], because level
splittings such as ±|T | depend on all other electrons in the system.
An alternative view on the Stoner limit is that antisymmetrized wave functions
|Ψ > diagonalize the leading one-electron part (T -part) of Eq. (11) and can therefore
be used to evaluate electron-electron interactions (U and JD ) in lowest-order
perturbation theory. The antisymmetric wave functions have the character of Slater
determinants if the spin is included. For example, Fig. 3 corresponds to
 
|FM > = σ (r1 ) σ ∗ (r2 ) − σ ∗ (r1 ) σ (r2 ) ↑ (1) ↑(2) (19)

and

|AFM > = σ (r1 ) σ (r2 ) (↑(1) ↑(2) −↑ (1) ↑(2)) (20)

In this method, known as the independent-electron or quantum-mechanical mean-


field approximation in solid-state physics and the molecular-orbital (MO) method in
chemistry, individual electrons move in an effective potential or “mean field” Veff (r)
created by all electrons in the system (Section “Itinerant Exchange”).

Correlations

The quantum-mechanical mean-field approach, which is the rationale behind the


local-density approximation to density-functional theory (LSDA DFT), has been
highly successful in magnetism, but some red flags indicate the need for a more
thorough analysis. For example, the mean-field result of Eq. (18) leads to the
prediction of positive (FM) exchange for JD = 0 so long as |T | < 14 U . In fact,
putting JD = 0 in the exact result of Eq. (17) shows that the exchange is always
62 R. Skomski

negative for JD = 0. Ferromagnetic coupling requires


  
U
|T | < JD + JD (21)
2

which is qualitatively different from |T | < 14 U .


The limitations of the quantum-mechanical mean-field approach are defined by
the treatment of the correlation hole. The correlation energy is defined [9] as the
difference between the correct many-electron energy and the corresponding one-
electron (independent-electron) energy obtained from a single Slater determinant
(Hartree-Fock determinant).
Consider |Ψ AFM > of Eq. (20), whose real-space part has the structure

1
|σ σ > = (|LL> + |LR> + |RL> + |RR>) (22)
2

This wave function has an ionic character of 50%, that is, the electrostatically
unfavorable configurations |LL> and |RR> provide half the weight. Since the
electrons equally occupy all two-electron states, Eq. (22) lacks a correlation hole.
The Coulomb penalty associated with the unfavorable ionic contribution leads to an
overestimation of the AFM energy and therefore to an overestimation of the trend
toward ferromagnetism.
In reality, electron correlations lead to a partial suppression of the |LL> and
|RR > occupancies, described by the mixing angle χ in Eq. (14). The Heisenberg
limit, Ψ + in Eq. (2) and χ = 0 in Eq. (14), has |Ψ AFM > ∼ |LR> + |RL>, which
corresponds to an ionic character of 0% and to a fully developed correlation
hole. The Heisenberg model is said to be overcorrelated, as opposed to the
undercorrelated independent-electron approach. An interesting approach is the use
of Coulson-Fischer wave functions, that is, of Slater determinants constructed not
from |L> and |R> but from combinations such as |L> + λ |R>, where λ ≈ |T |/U for
small hopping [10]. This unrestricted Hartree-Fock approximation contains a part
of the correlations at the expense of symmetry breaking in the Hamiltonian [9]. The
approximation is sufficient to reproduce the correct AFM wave function, Eq. (14),
for the H2 model of Eq. (11), but this finding cannot be generalized to arbitrary
many-electron systems. Near the equilibrium H-H bond length of about 0.74 Å, the
electrons are delocalized, described by Eq. (22) and λ = 1, but above 1.20 Å, the
electrons localize very rapidly and λ approaches zero.
Correlations primarily affect AFM spin configurations [6]. For example, the FM
wave function |σ σ ∗ > – |σ ∗ σ > = |LR> – |RL>, Eq. (13), is independent of T and
U and therefore unaffected by correlations. The reason for the absence of ionic
configurations in Eq. (13) is the Pauli principle, which creates the exchange hole
and forbids |LL > and |RR > occupancies with parallel spin. Correlations effects are
most important in half-filled bands, where ferromagnetism means that all bonding
and antibonding real-space orbitals are occupied by ↑ electrons and the net energy
2 Magnetic Exchange Interactions 63

gain due to interatomic hybridization is zero. Electrons (or holes) added to half-
filled bands do not suffer from this constraint and make ferromagnetism easier to
achieve.
Solid-state correlations are multifaceted and yield many more or less closely
related magnetic phenomena, such as spin-charge separation (Section “Antiferro-
magnetic Spin Chains”), wave-function entanglement, and the fractional quantum-
Hall effect (FQHE). The determination of correlations is demanding even- or
medium-sized molecules or clusters, because the number of configurations to be
considered increases exponentially with system size. For example, the complete
description of a single CH4 molecule (10 electrons) requires the diagonalization of
a matrix containing 43,758 × 43,758 determinants [9]. Some methods to describe
correlations [9–13] are microstate approaches, such as those in this chapter, self-
energy methods, the evaluation of matrix elements between Slater determinants
(known as the configuration interactions, CI), dynamical mean-field theory (DMFT)
[14, 15], and the Bethe ansatz [16, 17]. Unlike LSDA+U, the DMFT is a
true correlation approach, because the electrons keep their individuality and the
mean-field character refers to the spatial aspect of the correlations only. Some
other correlation approaches, such as the Hubbard model, are briefly discussed in
Section “Hubbard Model”.

Heisenberg Model

In the strongly correlated Heisenberg limit (U


T ), Eq. (17) becomes

2T2
J = JD − (24)
U

Putting U = ∞ yields J = JD and reproduces the naïve Heisenberg result of


Eq. (3). Expressions very similar to Eq. (24) can be derived for solids [3], but the
method is cumbersome and the resulting picture not very transparent. It is often
better to eliminate hopping terms and to consider spin Hamiltonians.
To replace the real-space wave functions (R and L) by spins (↑ and ↓), one
considers the full wave functions in the Heisenberg limit, namely, the AFM singlet

1
|AFM> = (|LR> + |RL>) (|↑↓> − |↓↑>) (25a)
2

and a FM triplet

1
|FM ↑↑> = √ (|LR> − |RL>) |↑↑> (25b)
2

1
|FM0> = (|LR> − |RL>) (|↑↓> + |↓↑>) (25c)
2
64 R. Skomski

1
|FM ↓↓> = √ (|LR> − |RL>) |↓↓> (25d)
2

The triplet (25b-d) reflects Sz = (−1, 0, +1) for S = 2 · ½ = 1 and is split by an


external magnetic field (Zeeman interaction).
In the Heisenberg model, one considers the spin part and implicitly understands
that the spins are located on neighboring atoms. The model involves spin operators
S = 12 σ, where σ is the vector formed by the Pauli matrices. The mathematical
direct-product identity
⎛ ⎞
1 0 0 0
⎜0 1 1 0⎟
σx ⊗ σx + σy ⊗ σy + σz ⊗ σz = ⎜
⎝0
⎟ (26)
1 −1 0⎠
0 0 0 1

reproduces the eigenfunctions and the singlet-tripletsplitting of Eq. (25), so that the 
Heisenberg Hamiltonian can be written as H=–2 J S x ⊗ S x +S y ⊗ S y +S z ⊗ S z ,
in vector notation, H = –2 J S 1 · S 2 . An alternative approach is to apply angular-
momentum algebra to S = S 1 + S 2 , using S 2 = S 1 2 + 2 S 1 · S 2 + S 2 2 and
S 1 2 = S 1 2 = 3/4, and exploiting that S 2 = S(S + 1) is equal to 2 (S = 1,
↑↑), and S 2 = 0 (S = 0, ↑↓). Considering atomic spins of arbitrary size S ≥ 1/2,
performing a lattice summation over all spin pairs (compare Fig. 1), and including
an external magnetic field, the Heisenberg Hamiltonian becomes
 
H = −2 Jij S i · S j − g μo μB S1 · H i (27)
i>j i

where the Jij are often treated as parameters. Solutions of the Heisenberg model
will be discussed in Sections “Spin Waves and Anisotropic Exchange”, “Antiferro-
magnetic Spin Chains”, and “Dimensionality Dependence of Quantum Antiferro-
magnetism”.
Some definitions of J involve a factor of 2, depending on whether the summation
is over all atoms (subscript ij) or only over pairs of atoms (subscript i > j). Even
opposite signs are sometimes chosen, using J > 0 and J < 0 for AFM and FM
interactions, respectively. The most common definition of J , used in Eq. (27), is
actually an exchange per electron, not per atom. The AFM-FM energy difference per
pair of atoms, E(Sz = 0) – E(Sz = 2S), is equal to 4 S (S + 1/2) J and diverges in
the classical limit (S = ∞). The divergence is removed by introducing renormalized
atomic exchange constants Jat = 2S 2 J or Jat = 2S (S + 1). In the classical limit
(S = ∞), Jat = Jat and H = −Jat s 1 · s 2 , where the unit vector s = S/S = M/Ms
describes the local magnetization direction. The classical energy splitting between
the ↑↑ and ↑↓ states, namely, ±Jat , is formally the same as that for S = 1/2, ±J .
 2
Biquadratic exchange, H = –B S · S , as well as other higher-order
terms, may arise for several reasons, for example, in T /U expansions of the full
Hamiltonian [3]. In the case of spin 1/2 interactions, they do not yield new physics,
2 Magnetic Exchange Interactions 65

 2  
because S · S = 3/16 − 1
2 S · S , but biquadratic exchange effects are
nonzero for S ≥ 1.

Hubbard Model

Completely ignoring the small direct exchange JD in equations such as (11, 12, 13,
14, 15, 16, 17, 18) leads to the Hubbard model. Generalized to solids, the Hubbard
Hamiltonian is
 
+ +
H = i,j Tij ĉi↑ ĉj↑ + ĉi↓ ĉ↓ + U i n̂i↑ n̂i↓ (28)

where n̂ = ĉ+ ĉ [18, 19]. In the Hubbard model, correlation effects are described by
the Coulomb interaction U . Equation (21) indicates that the Hubbard model does not
predict ferromagnetism in half-filled bands, but this argument cannot be generalized
to arbitrary bands and band fillings.
The bare Coulomb interaction is very high, about 20 eV for the iron-series
elements, but this value is reduced to about 4 eV due to intra-atomic correlations and
screening by conduction electrons. The screening (Sections “Itinerant Exchange”
and “Metallic Correlations and Kondo Effect”) depends on the crystal structure,
and eg orbitals tend to have slightly higher U values than t2g orbitals, so that U
varies somewhat for a given element. Table I shows typical U values for the three
transition-metal series [20]. Note that the effects of U are complemented by the
moderately strong intra-atomic exchange Jat , also listed in Table I. Approximate
values for U in some main-group elements are 8.0 eV (C), 3.1 eV (Ga), and 4.2 eV
(As). In rare earths, U is equal to and best obtained from the spectroscopic Slater-
Condon parameter F0 . It is of the order 10 eV and somewhat increases with number
of 4f electrons.
The Hubbard U yields a number of correlation effects. One of them is the
suppression of metallic conductivity for large values of U (Mott localization),
which reflects the splitting of metallic bands into upper and lower Hubbard bands
with opposite spin directions. The effect is very similar to the Coulson-Fischer

Table 1 Typical values of nv U Jat U Jat U Jat


screened Coulomb integrals
3 Sc 2.4 0.4 Y 1.7 0.3 Lu 1.5 0.3
(U ) and screened intra-atomic
exchange (Jat )(Jat ) 4 V 3.1 0.5 Zr 2.4 0.4 Hf 2.0 0.3
5 Ti 3.3 0.6 Nb 2.7 0.5 Ta 2.4 0.4
6 Cr 4.5 0.7 Mo 3.7 0.5 W 3.5 0.5
7 Mn 4.5 0.7 Tc 3.9 0.6 Re 3.7 0.5
8 Fe 3.9 0.7 Ru 4.2 0.6 Os 4.1 0.5
9 Co 4.4 0.8 Rh 4.0 0.6 Ir 3.8 0.5
10 Ni 4.0 0.8 Pd 3.8 0.6 Pt 3.6 0.5
11 Cu 5.7 0.8 Ag 4.8 0.6 Au 4.0 0.6
66 R. Skomski

Fig. 4 Hubbard interpretation of band gaps: (a) Mott-Hubbard insulator, (b) charge-transfer
insulator, (c) simple interpretation of Hubbard-Mott transition, and (d) refined Hubbard transition
involving a correlated metal phase known as the Brinkman-Rice (BR) phase

electron localization in the H2 molecule (Section “Correlations”). Some oxides are


antiferromagnetic Mott-Hubbard insulators, Fig. 4a, but many are charge-transfer
insulators [21], where the 2p-3d gap Δ is smaller than the Hubbard gap U (Fig. 4b)
and the transition to metallic behavior involves hopping between cation 3d and anion
2p states. The trend toward charge transfer behavior increases from early to late
transition metals and from oxides to halides.
In spite of the simplicity of Eq. (28), there have been no exact solutions for
the Hubbard model so far, except for a few special cases. Even the well-known
Hubbard band splitting (Fig. 4c) is a simplification. A detailed analysis, using
Gutzwiller wave functions [19] and dynamical mean-field theory (DMFT) [14],
yields a correlated-metal or Brinkman-Rice phase [22] with metallic quasiparticles
in the middle of the Hubbard gap (Fig. 4d). This quasiparticle peak is analogous to
impurity peaks near band edges, for example, in the gaps of semiconductors. The
difference is that the disorder responsible for the peak is not caused by impurity
atoms but by correlated electrons (and holes) randomly occupying lattice sites.

Specific Exchange Mechanisms

The involvement of Coulomb integral (U ), and exchange integral (JD ), and one-
electron level splitting (T ) is a common feature of exchange interactions, but the
interplay between these quantities varies greatly among magnetic solids.
2 Magnetic Exchange Interactions 67

Intra-Atomic Exchange

Atomic wave functions inside a given atom are orthogonal, so that the ferro-
magnetic exchange is not weakened by one-electron level splittings involving
hopping between different orbitals (T = 0). On the other hand, one-electron energy
differences between shells and subshells are typically large, several eV. In terms of
Fig. 3, these energy differences provide a forbiddingly one-electron level splitting.
Ferromagnetic intra-atomic exchange is therefore almost exclusively limited to
the nearly degenerate electrons in the partially filled inner subshells of transition-
metal atoms, namely, 3d, 4d, and 5d electrons in the iron, palladium, and platinum
series, respectively, 4f electrons in rare-earth (lanthanide) atoms, and 5f electrons
in actinides.

Hund’s Rules Intra-atomic exchange and spin-orbit coupling give rise to the
hierarchy of three Hund’s rules [23]. The rules, which are empirical but have a
sound physical basis, determine the magnetic ground state of atoms or ions. Hund’s
first rule reflects intra-atomic exchange and states that the total spin S is maximized
so long as the Pauli principle is not violated. The number of one-electron orbitals
per subshell is 2 l + 1, which yields 5 orbitals per d-shell and 7 orbitals per f-shell.
In the first half of each series, all spins are ↑, and for half-filled shells, the total spin
moment is therefore 5 μB (d-shells) and 7 μB (f -shells). Additional electrons are ↓
due to the Pauli principle. For example, Co2+ has a 3d7 electron configuration and
the spin structure 3d (↑↑↑↑↑↓↓).

Quantum states characterized by quantum numbers L and S form terms denoted


by 2S + 1 L. For example, the term symbol 2 F means L = 3 and S = ½. The next
consideration is Hund’s second rule, which states that the orbital angular moment
L is maximized, subject to the value of S. The vector model usually employed
in magnetism assumes L-S (Russell-Saunders) coupling, where the total orbital
moment L = i Li and the total spin moment S = i S i combine to yield the total
moment J = L + S. The operators obey angular-momentum quantum mechanics,
for example, S 2 = S (S + 1), L2 = L (L + 1), and J2 = J (J + 1). The opposite
limit of j-j coupling, where the spin-orbit interaction dismantles the total ionic spin
and orbital moments, is important only for the ground state of very heavy elements
(Z > 75) and for excited states of most elements [24], which are usually of no
concern in magnetism.
Spin-orbit coupling causes the terms to split into multiplets, which are denoted
by 2S + 1 LJ, and obey |L – S| ≤ J ≤ |L + S|. Hund’s third rule describes how spin (S)
and orbital moment (L) couple to yield the total angular momentum (J): Less than
half-filled subshells have J = |S – L|, whereas more than half-filled shells exhibit
J = |S + L|. This rule explains, for example, the large atomic magnetic moments of
the heavy rare earths, such as 10 μB per atom in Dy3+ and Ho3+ .
68 R. Skomski

Consider, for simplicity, the Hund’s-rules ground state of the p2 configuration,


realized, for example, in free carbon atoms. There are six one-electron states (px↑ ,
py↑ , pz↑ , px↓ , py↓ , pz↓ ), but the Pauli principle reduces the 6 × 6 = 36 two-electron
microstates to 15 Slater determinants. For example, |↑ ◦ ↓> ∼ |x↑ (r1 )y↓ (r2 ) –
y↓ (r1 )x↑ (r2 )> means that the px (Lz = +1) and py (Lz = −1) orbitals are both
occupied by ↑ electrons, while the pz orbital (Lz = 0) is empty. The 15 microstates
of the p2 configuration form three terms: 1 S (L = 0, S = 0), 1 D (L = 2, S = 0), and
3 P (L = 1, S = 1). Hund’s first rule uniquely establishes the ground-state term 3 P,

because the other two terms have zero spin. The term contains (2 L + 1) (2S + 1) = 9
Slater determinants, for example, |↑ ↑ ◦>, where Lz = 1 and Sz = 1. Hund’s third
rule predicts J = L – S = 0, corresponding to a nonmagnetic ground state. The p2
ground-state wave function is a superposition of three Slater determinants described
by Clebsch-Gordan coefficients C(L, Lz , S, Sz |J, Jz ) [2, 25]. Explicitly

1 1 1
|ψ> = √ |↑ ◦ ↓> + √ |↓ ◦ ↑> − √ |◦ ↑↓ ◦> (29)
3 3 3

The involvement of two or more Slater determinants indicates that correlations are
not necessarily be important even in seemingly simple systems.
Hund’s rules are obeyed fairly accurately by rare-earth ions in metallic and
nonmetallic environments. For example, the ground-state multiplets of rare-earth
ions obey J = |L ± S|, whereas excited multiplets have relatively high energies, with
notable exceptions of Eu3+ and Sm3+ , where the splitting is only about 0.1 eV [26].
One reason for the applicability of the rules is that the 4f -shell radii of about 0.5 Å
are much smaller than the atomic radii of about 1.8 Å. This enhances the spin-orbit
coupling and reduces the interaction with surrounding atoms. By contrast, Hund’s
rules are often violated in 3d, 4d, and 5d transition metals, where orbital moments
are quenched.

Moment Projections and Quenching Exchange interactions are between spins S,


not between total moments J=L+S, which makes it necessary to project the total
moment onto the spin moment. Similarly, the Zeeman interaction with an external
field involves L + 2 S, not J = L + S and S. In the Zeeman case, projection of
L + 2 S onto J yields the symbolic replacement L + 2 S → g and the moment
m = g J. The g-factor is obtained by using (L + 2 S) · J = g J2 and

J (J + 1) = L (L + 1) + 2L · S + S (S + 1) (30)

The result of the calculation is

3 1 S (S + 1) − L (L + 1)
g= + (31)
2 2 J (J + 1)

which yields g = 1 – S/(J + 1) and g = 1 + S/J for the first and second
halves of the lanthanide series, respectively. To account for spin-only character of
2 Magnetic Exchange Interactions 69

interatomic exchange, the atomic projection S · J = (g − 1) J2 must be used.


The corresponding de Gennes factor G = (g – 1)2 J(J + 1) is important for
the finite-temperature behavior of rare-earth magnets, where it controls the Curie
temperature. One implication of Eq. (31) is that the vectors L, S, and J are not
necessarily (anti)parallel but described by the vector model of angular momenta
[24]. A good example is Sm3+ , which has antiparallel spin and orbital moments
L = 5 and S = 5/2, respectively, so that L – 2S could naïvely be expected to yield
a zero magnetic moment. In fact, g = 2/7 and m = 0.71 μB , which corresponds to
angles of 22◦ between L and J and of 44◦ between S and J.
Hund’s rules are often violated in metallic and nonmetallic transition-metal
magnets. The d orbitals of iron-, palladium-, and platinum-series atoms are fairly
extended, so that interactions with neighboring atoms outweigh Hund’s rules
considerations. The rules regarding L are affected most, because the orbital moment
is normally quenched, L ≈ 0. For example, bcc iron has a magnetization of about
2.2 μB , but only about 5% is of orbital origin. The reason is that orbital moments
require an orbital motion of the electrons, but this motion is disrupted by the crystal
field introduced in Section “One-Electron Wave Functions”. Note that L = 0 means
J = S and, according to Eq. (31), g = 2.

High-spin Low-spin Transitions Crystal-field interactions cause the five 3d levels of


transition-metal ions to split. In magnets with cubic crystal structure, this splitting
is of the eg -t2g type: The |z2 > and |x2 – y2 > orbitals form the eg dublet, whereas
the |xy>, |xz>, and |yz> orbitals form the t2g triplet. The crystal-field interaction
yields a moment reduction if the splitting is larger than the combined effect of U
and JD . Such transitions are known as high-spin low-spin transitions. For example,
in octahedral environments, the energy of the t2g triplet is lower than that of the
eg dublet. Co2+ has seven 3d electrons, which translate into the spin configuration
t2g (↑↑↑↓↓) eg (↑↑) and a moment of 3 μB . However, in the limit of large crystal-
field splitting, one of the two eg↑ electrons “falls down” in the sense of Fig. 3 and
occupies the empty t2g↓ orbital, yielding the spin configuration t2g (↑↑↑↓↓↓) eg (↑)
and a moment of 1 μB . Examples are the Co2+ complexes [Co(H2 O)6 ]2+ (high
spin) and [Co(CN)6 ]4− (low spin).

Indirect Exchange

The model of Section “Electron-Electron Interactions” describes the so-called direct


exchange between nearest neighbors, where the hopping integral T competes
against U and JD . Exchange in solids is often indirect, mediated by conduction
electrons or by intermediate atoms, such as oxygen.

Superexchange Transition-metal oxides frequently exhibit exchange bonds of type


Mm+ -O2− -Mm+ , where Mm+ is a transition-metal cation. This type of exchange
is known as superexchange and also realized in magnetic halides such as MnF2 .
70 R. Skomski

The net exchange is tedious to calculate [27], but a transparent physical picture
emerges if one assumes that U and JD compete against T and that the outcome of
this competition is largely determined by T , similar to Eq. (24). For one 3d level
per transition-metal atom (M) and one oxygen 2p level (O), Eq. (8) becomes

⎛ ⎞
EM Tpd(L) 0
H = ⎝ Tpd(L) EO Tpd(R) ⎠ (32)
0 Tpd(R) EM

Here, EM and EO are the atomic on-site energies, and Tpd(R/L) describes the hopping
between M and O atoms. When Tpd(R) = Tpd(L) , a unitary transformation using

⎛ ⎞
√1 0 √1
2 2
⎜ ⎟
=⎝ 0 1 0 ⎠ (33)
√1 0 √1
2 2

partially diagonalizes the Hamiltonian and yields

⎛ √ ⎞
√EM 2Tpd 0
Q+ HQ = ⎝ 2Tpd EO 0 ⎠ (34)
0 0 EM

The transformation couples the wave functions of the two transition-metal atoms.
One of the coupled M levels is nonbonding (bottom-right matrix element), whereas
the other one (top left) hybridizes with the oxygen, thereby creating a level splitting
between the two coupled M orbitals. In the Heisenberg limit, Tpd is small, and the
diagonalization of Eq. (34) yields the transition-metal level splitting ±Teff , where
Teff = Tpd 2 /|EM –EO |.
Substitution of Teff into Eq. (24) yields the effective transition-metal exchange

2 Tpd 4
Jeff = JD − (35)
U (EM − EO )2

Since JD is small, the hopping normally wins, and the exchange in most oxides is
therefore antiferromagnetic. However, the non-s character of the 2p and 3d orbitals
causes Tpd to depend on the type of d orbital (eg or t2g ) and on the bond angle.
Figure 5 compares a 180◦ bond (a) with a 90◦ bond (b). In (a), the two p-d bonds
differ by the sign of the involved 2p wave-function lobe, but Tpd(R) = −Tpd(L)
leaves Eq. (35) unchanged. In (b), Tpd(R) = 0 by symmetry, because the hopping
contributions of the two oxygen lobes (+ and –) cancel each other. This implies
Tpd(R) = 0 in Eq. (32), and the two transition-metal orbitals are no longer coupled
(Teff = 0).
2 Magnetic Exchange Interactions 71

Fig. 5 Overlap and


exchange: (a) nonzero
overlap (180◦ bond) and zero
overlap (90◦ bond). In (a), the
hopping integral is nonzero,
corresponding to
antiferromagnetic indirect
exchange, but in (b), the
hopping integral is zero by
symmetry

The above analysis is the basis for the Goodenough-Kanamori-Anderson rule


[27, 28], which
 states that exchange in oxides is antiferromagnetic for bond angles
θ B > 90◦ Teff 2 > 0 but ferromagnetic for bond angles of θ B = 90◦ (Teff = 0).
Examples of the former are rock salt, spinel, and wurtzite oxides, where the
predominant bond angles are 180◦ , 125◦ , and 109◦ , respectively [27]. Ferromagnetic
exchange dominates in CrO2 [27], where the Cr4+ ions yield a net moment of 2 μB
per formula unit.

Ruderman-kittel Exchange The exchange interaction of localized magnetic


moments in metals is mediated by conduction electrons, which is known as the
Ruderman-Kittel-Kasuya-Yosida or RKKY mechanism. Electrons localized at Ri
and conduction electrons of wave vector k undergo a strong intra-atomic s-d
exchange –Jsd S k · S i δ (r–R i ), so that the localized electrons perturb
 the sea
of conduction electrons. The perturbed wave functions are ψ k (r) = k ck exp (i
k · r) dk, where the integration is limited to wave vectors |k| < kF . The wave-
vector cutoff affects the real-space resolution of the response ψ(r) and means
that details smaller than about 1/kF , such as δ(r – Ri ), cannot be resolved. As
a consequence, the electron density n(r)∼ ψ k (r) ψ k (r) dk contains a wave-
like oscillatory contribution. The oscillations are spin-dependent and yield the
oscillatory RKKY exchange
72 R. Skomski

2kF R cos (2kF R) − sin (2kF R)


J (R) = Jo (36)
(2kF R)4

between localized moments at Ri and Rj = Ri + R. In metals, kF ∼ n1/3 is large, and


the oscillation period does not exceed a few Å. In dilute magnetic semiconductors
(DMS), n can be made small by adjusting doping level and/or temperature, and the
RKKY interaction is then a nanoscale effect.
Equation (36) describes exchange interactions mediated by free electrons, but
the underlying perturbation theory can also be used to treat arbitrary independent-
electron systems, such as tight-binding electrons in metals [29] and DMS exchange
mediated by shallow nonmagnetic donors (or acceptors) [30]. At finite temperatures,
the thermal smearing of the Fermi surface yields an exponential decay of the
oscillations, with a decay length proportional to kF /T.

Double Exchange Intra-atomic exchange favors parallel spin alignment, and elec-
trons retain a “spin memory” while hopping between atoms. This process translates
into a ferromagnetic exchange contribution first recognized by Zener [28, 31].
Double exchange occurs in mixed valence oxides, such as Fe3 O4 . This oxide
contains Fe3+ and Fe2+ ions on B-sites. The latter can be considered as Fe3+ ions
plus an extra electron that can hop more or less freely between the d5 ion cores.

The double-exchange mechanism is important in magnetoresistive perovskites


(manganites). The parent compound, LaMnO3 , contains Mn3+ ions only and is
an antiferromagnetic insulator. Partially replacing La3+ by Sr2+ creates a charge
imbalance that is compensated by the formation of Mn4+ ions. In both Mn3+ and
Mn4+ , the low-lying t2g triplets are occupied by three well-localized 3d electrons,
but in Mn3+ , there is an additional eg electron that yields ferromagnetic double
exchange and metallic conductivity.

Itinerant Exchange

The magnetism of 3d, 4d, and 5d elements and alloys is fairly well described by the
independent-electron approximation, which corresponds to the use of a single big
Slater determinant. The electrons move in the solid, and the corresponding hopping
competes against the electrostatic electron-electron interaction.
The simplest approach is to replace the crystal potential V(r) by a charge-
neutralizing homogeneous background V(r) = const. (jellium model). The only
free parameter describing the corresponding homogeneous but not necessarily free
electron gas is the electron density n. It is convenient to parameterize n in terms
of the average inter-electronic distance rs = (3/4πn)1/3 and to relate rs to the free-
electron Fermi wave vector kF = (9π/4)1/3 /rs . Typical values of kF ao are 0.34 (Cs),
0.72 (Cu), and 1.03 (Be) [8]. The inverse magnetic susceptibility of the jellium
is [32]
2 Magnetic Exchange Interactions 73

χp π 1
=1− + 2 (0.507 ln (kF ao ) − 0.162) (37)
χ kF ao kF ao 2

where χ p = (α/2π)2 ao kF is the susceptibility of the non-interacting electron gas


(Pauli susceptibility). The onset of ferromagnetism corresponds to χ = ∞, that is,
to 1/χ = 0.
Equation (37) includes the key distinction between kinetic energy (hopping),
scaling as 1/rs 2 ∼ kF 2 , and Coulomb interaction, scaling as 1/rs ∼ kF . The Pauli
susceptibility reflects the kinetic energy, whereas –π/kF ao is the independent-
electron Coulomb correction, which corresponds to Bloch’s early theory of itinerant
exchange [8, 33]. As the electron gas gets less dense and kF becomes smaller, the
π/kF ao term in Eq. (37) predicts ferromagnetism for kF ao < 1/π, which is close
to the value for alkali metals such as Cs. Experimentally, the alkali metals are not
particularly close to ferromagnetism, which is caused by d and f electrons, not by a
homogeneous electron gas.
In fact, the last term in Eq. (37), which scales as 1/kF 2 and reflects the
so-called random-phase approximation (RPA), negates the Bloch prediction of
ferromagnetism – χ (kF ao ) never reaches zero in Eq. (37). The physics behind the
RPA is that the charge of any individual electron is screened by the other electrons
in the metal, which amounts to a reduction of the net Coulomb repulsion from 1/rs
to an exponentially decaying interaction. In other words, the screening electrons
form a quasi-particle cloud around the electron and renormalize the Coulomb
interaction.
The Stoner theory replaces Eq. (37) by the semiphenomenological expression

χp
= 1 − I D(EF ) (38)
χ

where the Stoner parameter I ∼ 1 eV [34] describes the electron-electron interaction


(Section “Stoner Limit”). Equation (38) predicts ferromagnetism for high densities
of states (DOS), when the paramagnetic state becomes unstable and the magnets
satisfy the Stoner criterion (EF ) > 1/I. The DOS of d electrons is much higher than
that of the jellium electrons implied in Eq. (37), which explains the occurrence
of ferromagnetism in transition metals. Alternatively, since the DOS (density of
states) is inversely proportional to the bandwidth W ∼ |T |, ferromagnetism
occurs in narrow bands. This finding is in agreement with the general analysis of
Section “Antiferromagnetic Spin Chains”.

Band Structure and Magnetism The hopping aspect of magnetism is determined by


the band structure and by the metallic density of states (DOS). Both are obtained
from the eigenvalues and eigenfunctions of Hamiltonians of the type

2 2
H=− ∇ + j Vo (r − r j ) (39)
2me
74 R. Skomski

where the lattice-periodic potential depends, in general, on the electron distribution.


The eigenfunctions of Eq. (39) are Bloch states ψ(r) = exp (ik · r) u(r) and electron
densities n(r) = u*(r)u(r). Equation (39) describes delocalized electrons whose
electrical conductivity is infinite due to the absence of scattering matrix elements.
This includes the tight-binding limit of well-separated atoms, where the hopping
integrals decrease exponentially with interatomic distance, T ∼ exp (–R/Ro ), but
the conductivity remains infinite even for large R [8]. At zero temperature, the
magnets are well described by these Bloch-periodic wave functions. This includes
the explanation of non-integer moments, which are caused by the smearing of one-
electron wave functions and spin densities over many lattice sites.

Inhomogeneous Magnetization States Wave-function and magnetization inhomo-


geneities may have several reasons. Wave-function localization requires the break-
ing of structural periodicity due to disorder (Anderson localization) or finite
temperature. Near Tc , atomic-scale itinerant moments behave like Heisenberg spin
vectors (“spin fluctuations”) of random orientation but well-conserved magnitude,
the latter involving some short-range order. Experimentally, this localization man-
ifests itself as a characteristic specific-heat contribution [9]. This spin-fluctuation
picture is realized both in strong ferromagnets (e.g., Co), where the ↑ band is filled,
and in weak ferromagnets such as Fe, where the ↑ band is only partially filled.
Deviations from wave-function periodicity also occur due to electron correlations
(Mott localization, Section “Hubbard Model”), competing exchange in perfectly
periodic lattices (Section “Magnetic Order and Noncollinearity”), and surface
effects.

Very weak itinerant ferromagnets (VWIFs), such as ZrZn2 (Tc = 17 K), barely
satisfy the Stoner criterion, and their behavior is qualitatively different from that
of strong and weak ferromagnets [35, 36]. Thermal excitations act as local spin
perturbations that can be described by the wave-vector-dependent susceptibility
χ (k) [3]. For VWIFs, a good approximation is

χo
χ= (40)
|I D (EF ) − 1 + f (k)|

and f (k) = a2 k2 . Here χ o is the interaction-free susceptibility, approximately equal


to the Pauli susceptibility χ p of Eq. (37), and a is an effective interatomic distance.
Inverse Fourier transform of Eq. (40) yields |M(r)| ∼ exp.(−r/ξ ), where r is the
distance from the perturbation and ξ = a/|1–I D (EF )|1/2 . In VWIFs, I D ≈ 1,
so that ξ is large by atomic standards and blurs the distinction between intra- and
interatomic exchange. The Stoner transition, I D = 1, yields ξ = ∞ and corresponds
to Bloch-periodic wave functions. A rough Curie temperature approximate is [37].

Tc 2 Tc
2
+ =1 (41)
TS TJ
2 Magnetic Exchange Interactions 75

This equation interpolates between the Heisenberg limit TJ (spin rotations) and the
Stoner limit Ts (moment reduction).
Strongly exchange-enhanced Pauli paramagnets, such as Pt, are close to the
onset of ferromagnetism and have I – 1/(EF )  0. Magnetic impurities create spin-
polarized clouds of radius ξ in these materials. The corresponding radial dependence
M(r) of the magnetization combines a pre-asymptotic exponential decay (r  ξ )
with RKKY oscillations for large distances (r  ξ ). The exponential decay length
ξ is described by Eq. (40), in close analogy to VWIFs. For example, magnetic
surfaces of Co2 Si nanoparticles spin-polarize the interior of the particles with a
penetration depth ξ [38]. Spin polarized clouds in strongly exchange-enhanced
Pauli paramagnets are also known as a paramagnons [3]. Left to themselves,
these quasiparticles slowly decay, and by considering the time dependence of the
fluctuations, f (k) → f (k, ω) in Eq. (4), one can show that the relaxation time diverges
at the phase transition (critical slowing down).

Bethe-Slater Curve

It is of practical importance to have some guidance concerning the strength and


sign of the exchange in a given metallic magnet. An early attempt was the
semiphenomenological Bethe-Slater-Néelcurve [39], which plots the net exchange
or the ordering temperature as a function of the interatomic distance or number
of electrons. There are many versions of this curve, and Fig. 6 shows one of
them. The curve predicts antiferromagnetism for small interatomic distances,
ferromagnetism for intermediate distances, and the absence of magnetic order in
the limit of very large distances. Experiment, the results of Section “Stoner Limit”,
and detailed calculations [40] grant some credibility to the approach, but the curve
has nevertheless severe flaws [27, 41].
Equation (38) shows that the onset of ferromagnetism is predominantly deter-
mined the density of states (EF ) at the Fermi level. This density somewhat increases

Fig. 6 Early version of the


Bethe-Slater-Néel curve
[27, 39]
76 R. Skomski

with interatomic distance, but a more important consideration is the position of


the Fermi level relative to the big peaks in the DOS. These peaks tend to vary
substantially among materials with similar chemical composition but different
crystal structures. For example, many transition-metal-rich intermetallic alloys have
interatomic distances of about 2.5 Å but show big differences in spin structures and
magnetic ordering temperatures.
A specific example is the distinction between fcc and bcc Fe structures. First, the
interatomic distance R = 2Rat in fcc iron, 2.53 Å, is actually a little bit larger than
that in bcc Fe, 1.48 Å, so that Fig. 6 cannot explain the ferromagnetism of bcc Fe.
Second, the plot ignores that bcc and fcc iron have very different crystal structures.
One difference is the number of nearest neighbors, namely, 8 in the bcc structure and
12 in the fcc structure. The bandwidth increases with the number z of neighbors, so
the ferromagnetism tends to be more difficult to create in dense-packed structures
(z = 12 . . . 14) and easier to create at surfaces (z = 4 . . . 6).
However, the number of neighbors is not the main consideration, because fcc Ni
and fcc Co have 12 nearest neighbors but are both ferromagnetic. More important is
the location of the big peaks in the density of states. For nearly half-filled d-shells
(Cr, Mn), one wants to have the peaks somewhere in the middle of the band, whereas
for nearly filled d-shells (Co, Ni), main peaks near the upper band edge are preferred.
An accurate determination of the peak positions  can only be done numerically, but
the moments theorem [42], dealing with µm = Em (E) dE, provides some guidance
[41]. The respective zeroth, first, and second moments describe the total number of
states, the band’s center of gravity, and the bandwidth, all unimportant in the present
context.
The third moment, μ3 , parameterizes the asymmetry of the DOS, that is, whether
the main peaks of the DOS are in the middle of the band (μ3 = 0) or close to
the upper band edge (μ3 < 0). It can be shown [42] that μ3 reflects the absence
or presence of equilateral nearest-neighbor triangles in the structure, the former
yielding centered main peaks and the latter creating main peaks near the upper
band edge. Figure 7 provides a very simple example of this relationship. Equilateral
nearest-neighbor triangles are present in the fcc structure but not in the bcc structure,
which corresponds to bcc ferromagnetism in the middle of the series and fcc
(or hcp) ferromagnetism for Co and Ni. Fe is intermediate, but bcc Fe becomes
ferromagnetic more easily than fcc Fe.

Manganese Isolated manganese atoms have half-filled 3d shells and a magnetic


moment of 5 μB per atom, which corresponds to a magnetization of approximately
5 T in dense-packed Mn structures. If this magnetization could be realized in a
ferromagnetic material, it would revolutionize technology, particularly since Mn is a
relatively inexpensive element. However, most Mn-based permanent magnets, such
as MnAl, MnBi, and Mn2 Ga, exhibit rather modest magnetizations of the order of
0.5 T [43]. The main reason for the low magnetization of Mn magnets is the half-
filled character of the Mn bands.
2 Magnetic Exchange Interactions 77

Fig. 7 Crystallographic
motifs and density of states:
(a) square and (b) equilateral
triangle. The density of states
is largest in the middle (a)
and near the top of the level
distribution (b). The atomic
orbitals (red) are of the
s-type, but in [42], it can be
seen that 3d electrons behave
similarly, and the present
figure can be generalized to
three-dimensional lattices

Fig. 8 Exchange interactions


in hypothetical simple-cubic
Mn [45]

Magnetizations as high as μo Ms = 3.2 T (3.25 μB per atom) have been reported


in thin-film Fe9 Co62 Mn29 deposited on MgO [44], where DFT calculations predict
2.90 μB per atom [45]. A traditional interpretation in terms of Fig. 6 is that
dilution by Fe and Co atoms enhances the average distance between Mn atoms. The
tetragonal structure of the Fe-Co-Mn alloy is loosely related to that of L21 -ordered
Mn2 YZ Heusler alloys, where the Mn atoms occupy a simple-cubic sublattice and
exhibit ferromagnetic exchange [46]. DFT calculations (Fig. 8) actually indicate
that the Mn-Mn exchange never becomes ferromagnetic. Furthermore, the example
of L10 -ordered MnAl shows that large Mn-Mn distances are not necessary for
78 R. Skomski

ferromagnetic exchange: The dense-packed Mn sheets in the (001) planes of MnAl,


which form a square lattice, exhibit a strong FM intra-layer exchange J [47]. This
underlines the crucial role of atomic neighborhoods.

Metallic Correlations and Kondo Effect

The situation in 3d metals is intermediate between the uncorrelated itinerant limit


(U /W = 0) and the strongly correlated Heisenberg limit, with U /W ratios of the
order of 0.5 [9]. For example, electron-electron interactions cause a bare electron
to become surrounded or “dressed” by other electrons, forming a quasiparticle of
finite lifetime, because electrons constantly enter and leave the dressing cloud. The
corresponding relaxation time τ ≈ / Im (Σ), where Σ is the self-energy, decreases
with increasing interaction strength. For metallic electrons of energy Ek , the lifetime
is approximately EF 3 /V2 (Ek – EF )2 [48], meaning that weak interactions and
vicinity to the Fermi surface yield well-defined and slowly decaying quasiparticles
which constitute a Fermi liquid.
As pointed out in Section “Correlations”, the independent-electron approxi-
mation involves a single Stater determinant and does not account for correlation
effects. The treatment of correlations requires several Slater determinants, such
as the two determinants of the model of Section “Electron-Electron Interac-
tions” and the three determinants forming the ground state of the p2 config-
uration (Section “Intra-Atomic Exchange”). An example of correlated many-
electron states is the Gutzwiller wave function |> = exp –η i n̂i↑ n̂i↓ |o >,
where the parameter η depends on U /W and the exponential term has the
effect of creating new Slater determinants from |Ψ o > [9, 19]. The Gutzwiller
method can be interpreted as a many-electron extension of the Coulson-Fischer
approach.
It is sometimes claimed or implied that density-functional theory becomes exact
if one goes beyond the local-density approximation and that LSDA+U approaches
account for correlations. This argumentation is questionable for several reasons.
First, density-functional theory provides the correct ground-state energy [12, 49]
if the density functional is known, but the exchange interaction is an energy
difference between the ferromagnetic and other spin configurations (AFM, PM)
and therefore involves excited states. Second, the density functional is not known
very well. The local-density approximation uses a potential inspired by and well
adapted to nearly homogeneous dense electron gases. The eigenfunctions used in
LSDA, known as Kohn-Sham (KS) orbitals, are pseudo-wave functions without a
well-defined quantum-mechanical meaning. They serve to determine the density
functional [49] and lack, for example, Gutzwiller-type projection features. The local
character of the LSDA, which can be improved by gradient corrections [50], is not
essential in this regard: Hartree-Fock theory involves a single Slater determinant
but is highly nonlocal [8]. Other density functionals, such as the Runge-Zwicknagel
functional for highly correlated electrons in dimers [51] and the density functional
2 Magnetic Exchange Interactions 79

for Bethe-type crystal-field interactions of rare-earth 4f electrons [52], bear little or


no resemblance to the LSDA.
The underlying problem is that the density functional is a generating functional
very similar to partition function Z and free energy F= – kB T ln Z in equilibrium
thermodynamics [52, 53]. The generating functionals correspond to Legendre
transformations, and in thermodynamics, the transformations are realized through
the term – T S, where S is the entropy. Once Z is determined by the summation
or integration over all microstates, such as the atomic positions ri in a liquid,
temperature-dependent physical properties are obtained in a straightforward way
from F(T). The theory is exact in principle, but the predictions depend on the
accuracy of the partition function. One example is that low- and high-temperature
expansions have different domains of applicability. Another example is the statis-
tical mean-field approximation, including Oguchi-type nonlocal corrections [54],
which are unable to describe critical fluctuations. In density-functional  theory, the
Legendre transformation is realized through the integral – V(r) n(r) dr [53].
The density functional is obtained by eliminating the microstate information in
Ψ ( . . . , ri , . . . , rj , . . . ) and yields the ground state for each lattice potential V(r).
This lattice potential is the DFT equivalent of the temperature in thermodynamics,
and the accuracy of the predictions depends on the quality of the generating
functional.
The density functionals used in LSDA are not calculated but obtained through
intelligent and experimentally supported guesswork. An exception is the weakly
correlated limit (U ≈ 0), where the KS orbitals become quantum-mechanical
wave functions with well-defined physical meaning. There are two reasons for
the great success of the LSDA, and its extensions have two main reasons. First,
transition metals are only weakly correlated and therefore amenable to ad hoc
improvements using “second-principle” approaches (materials-specific choices of
methods and parameters). Second, the KS Slater determinants used in LSDA are of
the unrestricted Hartree-Fock type (Section “Correlations”), which are constructed
from wave functions having symmetries lower than that of the Hamiltonian
[9, 10]. Unrestricted HF determinants can be expanded in terms of “regular” Slater
determinants and therefore contain some correlations [9].

The “Plus U” Method The LSDA+U method modifies the KS one-electron


potential by a potential that depends on the electron’s atomic orbital i, essentially
[55].

 
1
Vi (r) = VLSDA (r) + U − ni (42)
2

A crude approximation is U ∼ U . The presence of U suppresses ↑↓ occupancies


in highly correlated 3d and 4f orbitals. The LSDA+U can be used, for example,
to adjust the charge state of magnetic ions (configurations) to their experimental
values. Such adjustments are sometimes necessary, because there is only one
80 R. Skomski

Fig. 9 LSDA+U for bcc Fe: (a) magnetic moment, (b) weak ferromagnetism and (c) strong
ferromagnetism. The direct exchange and double-counting corrections are ignored in this figure

KS determinant available to account for the uncorrelated subsystem (one Slater


determinant) and for the ion’s intra-atomic couplings (several Slater determinants).
Strictly speaking, U is a well-defined first-principle quantity [55], not a fitting
parameter that can be chosen to obtain a desired computational result. Figure 9
illustrates this point for the magnetic moment of bcc Fe, calculated using the VASP
code for with U varying from 0 to 6. The moment m per Fe atom (a) exhibits an
increase from 2.21 μB to 3.07 μB , the experimental value being about 2.22 μB . Near
U = 0.9 eV (dashed vertical line), the slope dm/dU changes from about 0.4 μB /eV to
0.1 μB /eV, caused by the unphysical transition from weak to strong ferromagnetism
(b-c).
2 Magnetic Exchange Interactions 81

Fig. 10 Model describing quantum-spin-liquid corrections in solids. The quantum-mechanical


mean-field (MF) approximation self-consistently treats an independent electron in a sea of
surrounding electrons (gray) and corresponds to one Slater determinant. Atomic Heisenberg spins
having S = 1/2 yield 2z + 1 Slater determinants

Noncollinear Density-functional Theory The Heisenberg model is based on quan-


tum rotations of atomic spins of fixed magnitude S 2 = S (S + 1). This is a rough
approximation for transition metals, where electrons are delocalized (itinerant)
and atomic moments are often non-integer. However, rotations of electron spins
(S = 1/2), which are realized through Pauli matrices and yield spin-wave functions
such as ψ(θ ) = (cos½θ , sin½θ ), can be implemented in the LSDA and used to
describe noncollinear spin states, including antiferromagnets [56]. This approach
corrects, for example, much of the great overestimation of the Curie temperature in
the Stoner theory.

The spin-wave functions ψ(θ ) are of the quantum-mechanical mean-field type,


weakly correlated, and not eigenstates of the Heisenberg Hamiltonian. Figure 10
illustrates the many-electron aspect of the approximation. The model treats one
↓ electron in a sea of ↑ electrons. The left part of the figure corresponds to the
quantum-mechanical mean-field approximation, where electrons interact with an
effective medium. In a slightly more realistic picture, the interaction with z nearest
neighbors is individualized through exchange bonds, as shown for z = 3 and
z = 5.
The model, which assumes S = 1/2 Heisenberg spins and nearest-neighbor
exchange J< 0, is exactly solvable. The ground state has one ↑ electron and
z ↓ electrons, which leads to the involvement of (z + 1) Slater determinants. The
admixture of these determinants describes whether the ↓ electron stays in its original
central place (Néel state) or “leaks” into the crystalline environment [57]. The cal-
culation shows that the reversed spin occupies neighboring atoms (dark gray) with a
combined weight of 50%, thereby affecting net exchange and ordering temperature.
Sections “Antiferromagnetic Spin Chains” and “Dimensionality Dependence of
Quantum Antiferromagnetism” considers the lattice aspect of this spin leakage.

Exchange in the Kondo Model The Kondo effect, characterized by a resistance


minimum, is a correlation effect caused by the exchange interaction of localized
82 R. Skomski

Table 2 Kondo Cr Mn Fe Co
temperatures (in kelvin) for
Rh – 10 50 1000
some transition-metal
impurities in nonmagnetic Pd 100 0.01 0.02 0.1
hosts (gray column) [59, 65] Pt 200 0.1 0.3 1
Cu 1.0 0.01 22 2000
Ag 0.2 0.04 3 –
Au 0.01 0.01 0.3 200
Zn 3 1.0 90 –
Al 1200 530 5000 –

impurity spins with conduction electrons [9, 58]. Below the Kondo temperature
TK , the interaction couples the conduction electrons to the impurity spins, which
enhances the electrical resistivity. Some Kondo temperatures for Cr, Mn, Fe and Co
in various matrices are shown in Table 2.

The simplest Kondo model is of the Anderson-impurity type, where a single


conduction electron, described by a delocalized orbital |c>, interacts with a localized
state |f > [9]. The Coulomb U is negligibly small for the delocalized orbital |c> but
large for the localized orbital |f >. Furthermore, the on-site energy of the localized
electron (bound state) is lower than that of the delocalized electron by E. In terms
of the wave functions |ff >, |fc>, |cf>, and |cc>, the Hamiltonian is

⎛ ⎞
U − E T T 0
⎜ T 0 0 T ⎟
H=⎜

⎟ (43)
T 0 0 T ⎠
0 T T E

Since U
T , the |ff > state (energy U − E) does not play any role in the ground-
state determination. In the absence of hybridization (T = 0), the ground state would
be degenerate, |f c> ± |c f >, both states having the energy E = 0 and containing
one localized and one delocalized electron. The first excited antiferromagnetic state,
|cc> = |c↑ c↓ >, has the energy E = E, meaning that the localized electron becomes
a conduction electron.
The hopping integral T does not affect the ferromagnetic state |f c>−|c f >,
because a localized ↑ electron cannot hop into a delocalized orbital that already
contains a ↑ electron. However, the localized ↑ electron can hop into a delocalized
orbital containing an electron of opposite spin, which lowers the energy of the
antiferromagnetic state. The corresponding singlet (↑↓) ground state has an energy
of –2T 2 /E = –2JK , roughly translating into a Kondo-temperature of TK =
2T 2 /kB E. Above TK , the |↑↓> and |↑↑> states are populated with approximately
equal probability, the two electrons effectively decouple, and the resistivity drops.
In reality, there are many conduction electrons, so an integration over all k-states
is necessary [58]. The main contribution comes from electrons near the Fermi level,
2 Magnetic Exchange Interactions 83

which form a Kondo screening cloud of size ξ proportional to 1/TK and yield a
Kondo temperature TK = (W/kB ) exp (−1/ [2 JK D(EF )]). Due to its exponential
dependence on JK and (EF ), TK varies greatly among systems [59]. Table II shows
some examples. TK is lowest for impurities in the middle of the 3d series and largest
for nearly filled or nearly empty 3d shells, as exemplified by TK = 5000 K for
Ni in Cu. The dependence JK ∼ 1/S indicates that Kondo exchange become
less effective in the classical limit. Heavy-fermion compounds, such as UPt3 and
CeAl2 , can be considered as Kondo lattices where conduction electrons interact
with localized 4f or 5f electrons [9].

Exchange and Spin Structure

Exchange affects spin structure and magnetic order in many ways. It determines
the ordering temperature, gives rise to a variety of collinear and noncollinear
spin structures, and influences micromagnetism through the exchange stiffness
A. Exchange phenomena include quantum-spin-liquid behavior, high-temperature
superconductivity, and Dzyaloshinski-Moriya interactions.

Curie Temperature

In spite of its simplicity, the Heisenberg model (27) is very difficult to solve,
especially in two and three dimensions. A great simplification is obtained by using
the identity

S i · S J = S i · <S j > + <S i > · S j + Cij + co (44)

 the thermodynamic correlation term Cij = (S i − <S i >) ·


and neglecting
S j − <S j > and the constant co = <S i > · <S j >. The latter is physically
unimportant, because it does not affect the thermodynamic averaging. The former
is important only in the immediate vicinity of the Curie temperature, where it
describes critical fluctuations [60, 61]. Substituting Eq. (44) into Eq. (27) and
assuming z nearest neighbors of spin moments S yield the factorized single-spin
Hamiltonian

H = −2 z J S · <S> − 2μo μB S · H (45)

This equation amounts to the introduction of a mean field μo μB H =2 z J <S> and


maps the complicated Curie-temperature problem onto the much simpler problem
of a spin S in a magnetic field. This approximation (45) is the thermodynamic mean-
field approximation, which must be distinguished from the quantum-mechanical
mean-field approximation used to treat electron-electron interactions.
The partition function belonging to Eq. (45) is a sum over the 2 S + 1 Zeeman
levels Sz . The field dependence of <S> has the form of a Brillouin function (BS ),
84 R. Skomski

and self-consistently evaluating <S> yields the Curie temperature

2 S(S + 1)
Tc = zJ (46)
3 kB

A generalization of this equation to two or more sublattices will be discussed


in Section “Dimensionality Dependence of Quantum Antiferromagnetism”. This
generalization includes the Néel temperature of antiferromagnets.
The spin excitations leading to Eq. (46) consist in the switching of individual
spins S i . The corresponding energies are rather high, with temperature equiv-
alents close to Tc . At low temperatures, the mean-field approximation predicts
exponentially small deviations from the zero-temperature magnetization M(0),
which is at odds with experiment. In fact, the low-temperature behavior M(0) –
M(T) of Heisenberg magnets is governed by low-lying excitations (spin waves)
(Section “Spin Waves and Anisotropic Exchange”) and described by Bloch’s law,
M(0) – M(T) ∼ T3/2 , in three dimensions.

Magnetic Order and Noncollinearity

Depending on the sign of the interatomic exchange, there are several types of mag-
netic order. Figure 11 shows some examples. Often there are two or more sublattices
[4, 54], and the division into sublattices can be of structural or magnetic origin.
Ferrimagnetism (FiM) normally reflects chemically different sublattices, such as
Fe and Dy sublattices in Dy2 Fe14 B. Antiferromagnetism (AFM) is also caused by
negative interatomic exchange constants, but the different sublattices are chemically
and crystallographically equivalent. For example, CoO crystallizes in the rock-salt
structure, but the Co forms two sublattices of equal and opposite magnetization.
Ferromagnetism is frequently encountered in metals (Fe, Co, Ni) and alloys (PtCo,
SmCo5 , Nd2 Fe14 B), the latter having different ferromagnetic sublattices. CrO2 is a
ferromagnet, but most oxides and halides are antiferromagnetic (MnO, NiO, MnF2 )
or ferrimagnetic (Fe3 O4 , BaFe12 O19 ).
Many oxides of chemical composition MFe2 O4 crystallize in the spinel struc-
ture, which contains one cation per formula unit on tetrahedral sites [...] (M2+ ,
sublattice A) and two cations per formula unit on octahedral sites {...} (Fe3+ ,
sublattice B). The exchange between the A and B sublattices is negative, which
yields a ferrimagnetic spin structure. The cation distribution over the A and B sites
depends on both chemical composition and magnet processing. For example, Fe3 O4
crystallizes in the so-called “inverse” spinel structure [Fe3+ ] {Fe2+ Fe3+ }(O2− )4
[65]. The total magnetization, measured in μB per formula unit, is therefore
[−5] + {5 + 4} = 4.
In the classical limit, the mean-field Curie temperature is given by the lowest
eigenvalue of the N × N matrix in the equation

kB T <si > = j Jij <sj > (47)


2 Magnetic Exchange Interactions 85

Fig. 11 Spin structures (schematic): ferromagnets (FM), antiferromagnet (AFM), ferrimagnet


(FI), Pauli paramagnet (PM), and noncollinear spin structure (NC)

This matrix equation is easily generalized to quantum-mechanical case, by carefully


counting neighbors and using the appropriate de Gennes factors and Brillouin
functions [54, 62]. The number N of sublattices is equal to the number of
nonequivalent atoms. In disordered solids, all atom are nonequivalent and N → ∞.
For two sublattices A and B, Eq. (47) becomes

3 kB T <sA > = j JAA <sB > + JAB <sB > (48a)

3 kB T <sB > = j JBA <sB > + JBB <sB > (48b)

Here JAA/BB and JAB/BA are the classical intra- and intersublattice exchange
constants, respectively, and the factor 3 reflects the classical limit of the Brillouin
functions. The solution of Eq. (48) is
  
1
Tc = (JAA + JBB ) ± (JAA − JBB )2 + 4 JAB JBA (49)
6 kB

The two sublattices often have different numbers of atoms, so that JAB = JAB
in general, but the two intersublattice exchange constants enter Eq. (49) in the
form of the product JAB JBA , and it is sufficient to consider J ∗ = (JAB JBA )1/2 .
For one-sublattice ferromagnets, where JBB = J ∗ = 0, Tc is equal to JAA /3kB .
86 R. Skomski

Various scenarios exist for two-sublattice magnets. In the simplest AFM case, the
two intrasublattice exchange interactions JAA = JBB = 0 and J ∗ <0, yielding the
Néel temperature TN = –Tc = |J ∗ |/3kB .

Metallic Sublattices Sublattice effect also occur in metals. In rare-earth transition-


metal (RE-TM) magnets, the RE-TM exchange (spin-spin) coupling is AFM for
the light rare earths and FM for the heavy rare earths. The orbital moment of the
TM sublattice is negligible, but inside each rare-earth atom, L and S are antiparallel
for light RE and parallel for heavy RE. This yields the spin structures [S↑ ]TM [S↓
L↑ ]RE for the light and [S↑ ]TM [S↓ L↓ ]RE for the heavy rare earths. The large but
opposite moment of the heavy rare-earth atoms yields a zero net magnetization
in some transition-metal-rich alloys. This spin state is referred to as compensated
ferrimagnetism (CFiM) and normally occurs at some compensation temperature
T0 , because different sublattices tend to have different temperature dependences
of magnetization [54, 63]. This is one of the features that distinguish CFiM from
AFM. Compensation occurs quite frequently in ferrimagnets, including oxides such
as rare-earth garnets R3 Fe5 O12 .

A rule of thumb for the exchange in transition-metal alloys is the switch rule:
The exchange is negative for interactions between late and early transition-metal
atoms but positive otherwise. For example, Co and Pt are both late transition-metal
elements, so the Co and Pt moments in CoPt are parallel. While the switch rule
includes alloys containing heavy transition-metal atoms (3d-4d and 3d-5d alloys),
it is not very reliable for elements in the middle of the series [64]. It also describes
impurities in host lattices and RE-TM intermetallic compounds, because rare earths
count as early transition metals, with one 5d electron contributing to the exchange
[65].
With the exception of very weak itinerant ferromagnets, intersublattice inter-
actions in metals are well described by the Heisenberg model, and equations
like (49) provide good estimates of the ordering temperature [62]. For example,
transition-metal-rich rare-earth permanent magnets have JTT
J ∗ ≈ JRT and
JRR ≈ 0, so that the rare-earth contribution
 to the Curie temperature is given by
Tc ≈ (JTT /3kB ) 1 + J ∗2 /JTT 2 . As a function of the number of 4f electrons, it
peaks in the middle of the lanthanide series, because J ∗2 involves the de Gennes
factor (Section “Intra-Atomic Exchange”)

Noncollinear Spin Structures There is a rich variety of noncollinear spin structures.


Spin glasses are disordered materials whose local magnetization is frozen below
some spin-glass transition temperature Tf [66, 67]. The definition of Tf , the spin state
below and above Tf , the nature of the transition, and the microscopic description are
nontrivial, but there is normally a distribution of exchange interactions Jij , caused,
for example, by RKKY interactions between localized moments in a metallic host.
In the simplest case, Jij = ±Jo for the interaction with z neighbors. For z → ∞,
the eigenvalue distribution of the random matrix Jij obeys Wigner’s semicircle law

[66], and the corresponding mean-field estimate is Tf = z Jo /kB . The situation is
2 Magnetic Exchange Interactions 87

further complicated by different types of disorder that can occur. Chemical disorder
means atomic substitutions with little or no changes in atomic positions. Bond and
topological
  disorders involve substantial changes in atomic positions and in Jij =
|r i − r j | , but in the latter case, there is no continuous transformation connecting
the ordered and disordered lattices [67].

Helimagnetism arises when competing exchange interactions between nearest


and next-nearest neighbors yield spin spirals of wave vector k. Such structures
are realized, for example, in the heavy rare-earth elements, where k || ez [68].
Consideration of a-b planes labeled by magnetization angles θ n yields the classical
Heisenberg energy

E = –J n cos (θn+1 –θn ) –J n cos (θn+2 –θn ) (50)

where J and J are the exchange interaction between neighboring layers, respec-
tively. The energy (50) is minimized by the ansatz θ n + 1 = θ n + δ, where δ ∼ 1/k
is the magnetization rotation between subsequent layers:

 
J + 4J cos δ sin δ = 0 (51)

Aside from including FM (δ = 0) and AFM (δ = π) states,  this equation


 has
noncollinear or helimagnetic (0 < δ < π) solutions, δ = arccos −J /4J .
Noncollinear spins structures are very common for elements in the middle
of the iron series, notably Cr and Mn, where the exchange contains competing
antiferromagnetic interactions. Elemental Cr forms a spin-density wave where the
AFM sublattice magnetization exhibits a real-space oscillation with a periodicity of
about 6 nm [59].

Dzyaloshinski-Moriya Interactions Noncollinearity may also arise from relativistic


Dzyaloshinski-Moriya (DM) interactions [69–71], which occur in structures with
violated or “broken” inversion symmetry. Examples are MnSi [72], α-Fe2 O3
(hematite) [65], and structurally disordered magnets such as spin glasses [66], as
well as in artificial magnetic nanostructures [73]. DM interactions are described by
the Hamiltonian HDM = – i>j Dij · Si × Sj , wherethe direction  of the DM vector
Dij = −Dj i is given by Dij ∼ n (r i − r n ) × r j − r n . In this expression, i
and j denote the two DM-interacting spins, and rn is the position of a magnetic or
nonmagnetic neighbor (Fig. 12). Physically, d electrons hop from atom i to atom
n and then to atom j. Unless rn is located on the line connecting ri and rj (and the
cross product determining D is zero), the hopping sequence involves a change of
direction at rn , which creates a partial orbit around rn and some spin-orbit coupling
that affects the spins i and j. The DM interaction changes the spin projections onto
the plane created by the vectors ri – rn and rj – rn : It tries to make Si and Sj parallel
to ri – rn and rj – rn , respectively.
88 R. Skomski

Fig. 12 Dzyaloshinski-Moriya Interactions in (a–b) crystals and (b–c) thin films. The red atoms
are magnetic, whereas the blue and white atoms are nonmagnetic but have weak (white) and strong
(blue) spin-orbit coupling

Since DM interactions are caused by spin-orbit coupling, they are a weak


relativistic effect, comparable to micromagnetic dipolar interactions and to magne-
tocrystalline anisotropy. They compete against the dominant Heisenberg exchange
and create canting angles of the order of 1◦ in typical magnetic materials [74].
By contrast, noncollinearities due to competing ferromagnetic or antiferromagnetic
exchange (Eq. (51)) can assume any value between 0◦ and 180◦ . However, DM
canting angles substantially larger than 1◦ are possible in materials with weak
Heisenberg exchange (low Tc ).
DM effects are strongly point-group-dependent, and the absence of inversion
symmetry is a necessary but not sufficient condition [75]. For example, inverse cubic
Heusler alloys have zero net DM interactions in spite of their noncentrosymmetric
point group Td . Figure 12(a–b) illustrates DM interactions in an orthorhombic bulk
2 Magnetic Exchange Interactions 89

crystal without inversion symmetry (point group C2v ). The fictitious crystal has
an equiatomic MT composition, where M is a magnetic or nonmagnetic metallic
element and T is a transition-metal element. The structure yields a spin spiral in the
x-z plane, that is, perpendicular to the net DM vector. B20-ordered cubic crystals
such as MnSi (point group T) are unique in the sense that their space group (P21 3)
is achiral due to the 180◦ character of the 21 screw axis but becomes chiral through
the incorporation of a chiral MnSi motifs.
Figure 12(c–d) shows the effect of Dzyaloshinski-Moriya interactions in thin
films with perpendicular anisotropy and fourfold (C4v ) or sixfold (C6v ) symmetry
(side view). When a patch of magnetic material is deposited on a material with
strong spin-orbit coupling, for example, Co on Pt, the modified spin structure is
reminiscent of a hedgehog. Such DM interactions are of interest in the context of
magnetic skyrmions. For example, bubble domains in thin films have a nonzero
skyrmion number and therefore yield a topological Hall effect (THE) [76], but DM
interactions change the spin structure of the bubble and the THE, thereby adding
new physics.

Spin Waves and Anisotropic Exchange

The low-lying excitations in Heisenberg magnets are of the spin-wave or magnon


type. Spin waves are of interest in experimental and theoretical physics and also
important in applied physics (microwave resonance, exchange stiffness). Chapter
SPW is devoted to spin waves, and in this chapter, the focus is on exchange in spin-
1/2 Heisenberg magnets, where quantum effects are most pronounced. To solve the
ferromagnetic Heisenberg model, it is convenient to rewrite the exchange term in
Eq. (27) as

1 + − + 
S i · S i+1 = Si Si+1 + Si− Si+1 + Sz,i Sz,i+1 (52)
2

The Sz operators measure the spin projections, Sz |↑>= + 12 |↑> and Sz |↓>=
− 12 |↓>, but leave the wave function unchanged. The spin-flip operators S + and S −
rise and lower the spin by one unit, respectively: S + |↓>=|↑> and S − |↑>=|↓>.
Since the S = 1/2 Heisenberg model has only two spin states, S + |↑>= 0 and
S − |↓>= 0, or symbolically S + S + = 0 and S − S − = 0. The products of
S + and S − in Eq. (52) have the effect of interchanging spins of opposite sign:
|↑↓ > becomes |↓↑ > and vice versa.
The ferromagnetic state, symbolically |0 > = |↑↑↑↑↑↑↑↑...>, is an eigenstate
of the Hamiltonian, because each of the spin-flip terms contains an S + operator
that creates a zero. One might naively expect that a single switched spin creates an
excited eigenstate, for example, |i > = |↑↑↑↓↑↑↑↑...>, where Ri is the position
of the flipped spin. However, the spin-flip operators move the flipped spin and
thereby create wave functions |i + 1 > and |i–1>. The low-lying eigenstates
of the ferromagnetic chain are actually plane-wave superpositions of single-spin
90 R. Skomski

Fig. 13 Spin wave (schematic)

flips, |ψ k > = exp.(ik·Rj ) |j>. These wave-like excited states are the spin waves
or magnons. Each magnon corresponds to one switched spin, but the reversal is
delocalized rather than confined to a single atom (Fig. 13). The corresponding
excitation energy is E = 4(1 – cos(k a)). For arbitrary crystals and spins [63]
   
E(k) = 2S j J R j 1– cos k · R j (53)

where Rj is the distance between the exchange-interacting atoms.


Of particular interest is the long-wavelength limit, where the dispersion relation
(53) becomes quadratic. The three monatomic cubic lattices (sc, bcc, fcc) have [63]

E(k) = 2SJ a 2 k 2 (54)

This equation cannot be generalized to more complicated cubic crystals, because


E(k) is governed by the interatomic distance Rj , not by the lattice constant a,
which can be very large due to superlattice formation. Application of Eq. (53) to
crystals without second-order structural anisotropy and z nearest neighbors (distance
R) yields E(k) = 2zSJ (1– sin(kR)/kR), which has the long-wavelength limit

z
E(k) = SJ R 2 k 2 (55)
3

For sc, bcc, and fcc lattices, Eq. (55) is equivalent to Eq. (54). In good approxima-
tion, it can also be applied to slightly noncubic structures. For example, elemental
cobalt has R = 2RCo , where RCo = 1.25 Å nm is the atomic radius of fcc and hcp Co.
Strongly anisotropic structures, such as multilayers, require an explicit evaluation of
Eq. (53).
It is common to write this relation as E = D k2 , where D is the spin-wave
stiffness. In micromagnetism, it is convenient to write the exchange energy as

E = A [∇s]2 dV (56)

where A is the exchange stiffness. Comparison of Eqs. (54) and (56) yields
2 Magnetic Exchange Interactions 91

Table 3 Spin-wave stiffness Material D A


D and exchange stiffness A
meV/nm2 pJ/m
for some materials [78]
Fe 2.8 20
Co 4.5 28
Ni 4.0 8
Ni80 Fe20 2.5 10
Co2 MnSn 2.0 6
Fe3 O4 5.0 7

J
A = 2 c S2 (57)
a

where c is the number of atoms per unit cell (c = 1 for sc, c = 2 for bcc, c = 4 for
fcc). Similar to Eq. (54), Eq. (57) cannot be used for arbitrary crystals, whose lattice
constants can be very large, and for dilute magnets [90]. In terms of the interatomic
distance, the rule of thumb is A ≈ zS 2 J /5R. Values of spin wave stiffness D and
exchange stiffness A for some common magnets are given in Table 3.

Anisotropic Exchange Anisotropic exchange is a vague term, used for a variety of


physically very different phenomena, sometimes even for the Dzyaloshinski-Moriya
interactions.

Spin waves are affected by magnetocrystalline anisotropy, especially in noncubic


magnets. The anisotropy adds a spin-wave gap Eg = E(k = 0) to Eq. (54) and
also affects the exchange stiffness. For example, in uniaxial (tetragonal, hexagonal,
trigonal) magnets, one needs to distinguish A|| (along the c-axis) and A⊥ (in the a-
b-plane). The difference is particularly large in multilayers, where the intra-layer
exchange (A⊥ ) is often much stronger than the interlayer exchange (A|| ). The
Heisenberg interaction behind this type of anisotropic exchange remains isotropic,
as in Eq. (27), and the difference
 between
 A|| and A⊥ is caused by the nonrelativistic
bond anisotropy, Jij = J R i − R j [79].
Very different physics are involved in the so-called anisotropic Heisenberg
model, which derives from the (isotropic) Heisenberg model by the replacement

 
J ŝ · ŝ → J ŝx · ŝ x + ŝy · ŝ y + Jz ŝz · ŝ z (58)

 
The exchange anisotropy J = Jz –J /Jz , which has the same relativistic
origin as magnetocrystalline anisotropy and the Dzyaloshinski-Moriya interaction
(Section “Magnetic Order and Noncollinearity”),
 is normally
 very small compared
to the average or “isotropic” exchange Jo = 2J + Jz /3. However, J becomes
non-negligible when Jo is very small, for example, in some compounds with low
Curie temperature [80].
92 R. Skomski

The XY and Ising models are obtained by putting Jz = 0 and J = 0 in Eq.


(58), respectively. In classical statistical mechanics, these models have the spin
dimensions n = 2 (XY) and n = 1 (Ising), as compared to n = 3 (Heisenberg model),
n = ∞ (spherical model), and n = 0 [4, 60]. The spin dimension has a profound
effect on the onset of ferromagnetism in D-dimensional crystals. Ising ferromagnets
have Tc = 0 in one real-space dimension (D = 1) but Tc > 0 for D ≥ 2. Heisenberg
magnets have Tc = 0 in one and two real-space dimensions but Tc > 0 for D ≥ 3. For
all ferromagnetic spin dimensionalities, statistical mean-field theory is qualitatively
correct in D > 4 real-space dimensions, with logarithmic corrections in D = 4. For
the geometrical meaning of D = 4, see Figs. 7.9, 7.10 in Ref. [4].
Two-dimensional magnets (D = 2) are particularly intriguing. The Heisenberg
model (n = 2) predicts Tc = 0, but adding an arbitrarily small amount of uniaxial
anisotropy to the Heisenberg model yields Tc > 0 [81]. This feature has recently
received renewed attention in the context of the two-dimensional van der Waals
(VdW) magnetism. The two-dimensional XY model (n = 2, D = 2) yields a
Thouless-Kosterlitz transition with a power-law decay of spin-spin correlations but
no long-range magnetic order.
There are actually two types of Ising models, characterized by similar Hamilto-
nians, J = 0 in Eq. (58), but different Hilbert spaces. Ising’s original model is de
facto a classical Heisenberg model with infinite magnetic anisotropy [82, 83], which
leads to two spin orientations, ↑ and ↓. The quantum-mechanical Ising model, also
known as the “spin-1/2 Ising model in a transverse field” [84–86], is physically very
different. For example, it allows states with <sx > = 0 and <sy > = 0, whereas the
idea of the (original) Ising model is to suppress such states, <sx > = <sy > = 0.
Exchange anisotropy or exchange bias in thin films means that a pinning layer
yields a horizontal hysteresis-loop shift in a free layer. The bias is realized through
FM or AFM exchange at the interface between the pinning and soft layers, but its
ultimate origin is the magnetocrystalline anisotropy of the pinning layers, which
is often an antiferromagnet. The situation is physically similar to the horizontal
and vertical hysteresis-loop shifts sometimes observed in hard-soft composites,
which are inner-loop effects. Micromagnetically, the exchange-energy density is not
confined to the atomic-scale interface but extends into the pinning and free layers,
so that the net interlayer exchange energy per film area is generally very different
from the atomic-scale interlayer exchange [73].

Experimental Methods

There are many methods to investigate exchange, directly and indirectly, some of
which are briefly mentioned here. Magnetic measurements are used to determine
Curie temperatures Tc ∼ J , from which exchange constants can be deduced. The
low-field magnetization of antiferromagnets is zero, but high fields tilt the AFM
sublattices and yield a small magnetization M(H ) ∼ H /J .
2 Magnetic Exchange Interactions 93

The exchange may also be deduced from the low-temperature M(T) curves,
because the Bloch law involves the exchange stiffness. Magnetic force and, to
a much lesser extent, anomalous magneto-optic microscopies used to investigate
magnetic domain structures, which contain implicit information about the exchange
stiffness. A direct method to probe exchange is magnetic resonance.
Neutron diffraction and, to a much lesser extent, X-ray diffraction (XRD) are
important methods to probe spin structure. The magnetic XRD signal is much
weaker than the neutron-diffraction signal. Interatomic exchange can be probed
by a variety of methods, such as X-ray magnetic dichroism, which also allows
a distinction of L and S contributions to the atomic moments. Electron-transport
measurements are frequently used to gauge and confirm exchange effects.

Antiferromagnetic Spin Chains

Spin waves are particularly intriguing in antiferromagnets, whose low-lying states


correspond to the highest excited states in the ferromagnetic case [8, 16, 17]. By
analogy with FM ground state, |↑↑↑↑↑↑↑↑...>, one could intuitively assume that
the AFM ground state is a superposition of the two quasi-classical Néel states

|AFM (1)>=| ↑↓↑↓↑↓↑↓ · · · > (59)

and

|AFM (2)>=| ↑↓↑↓↑↓↑↓ · · · > (60)

However, the spin-flip terms in Eq. (52) do not transform Eqs. (59) and (60) into
each other but create pairs of parallel spins (spinons), for example

S4 + S5− |↑↓↑↓↑↓↑↓ · · · >=|↑↓↑↑↓↓↑↓ · · · > (61)

Using the Néel states to evaluate Eq. (27) yields an AFM ground-state energy of
−0.5 J per atom, compared to the exact Bethe result of −2 J (ln 2 − 1/4) =
−0.886 J [8]. Systems with such complicated ground states are also known as
quantum spin liquids (QSL). The underlying physics is very similar to the spin
mixing discussed below Fig. 10 but now involves an infinite number of spins.
The derivation of the Heisenberg model, Section “Hubbard Model”, was based
on the neglect of interatomic hopping. Strictly speaking, this is meaningful only
when U is large and the band is half-filled. In more- or less-than-half-filled bands,
a fraction of the electrons can move almost freely. Such magnets have both charge
and spin degrees of freedom, and the corresponding extension of the Heisenberg
chain is known as the Tomonaga-Luttinger model or Luttinger liquid [87, 88]. A
typical wave function is |↑↓↑↑↓◦↑↓>, which contains one hole. The model has a
number of interesting features. For example, spin and charge excitations move with
different velocities, the former being slower, because spin excitations have lower
94 R. Skomski

energies δE = ω than charge excitations. This is an example of a correlation effect


known as spin-charge separation [88]. By contrast, in the itinerant limit, charge and
spin degrees are closely linked. Spin-charge separation is important in the Kondo
mechanism (where low-energy spin flips determine the resistivity) and in high-
temperature superconductivity (Section “Dimensionality Dependence of Quantum
Antiferromagnetism”).
The electron distribution n(E) of a Luttinger liquid is very different from a Fermi
liquid [88]. Weak correlations in metallic magnets create particle-hole quasiparticles
but leave the Fermi surface otherwise intact. Strong correlations, as in the Luttinger
liquid, completely destroy the Fermi surface, and n(E) becomes a smooth function.

Dimensionality Dependence of Quantum Antiferromagnetism

The Luttinger liquid is a typical one-dimensional effect: Arbitrarily small pertur-


bations of structural, thermal, or quantum-mechanical origin destroy long-range
magnetic order. Quantum-spin-liquid effect in higher dimensions are generally
less pervasive but not necessarily unimportant. In antiferromagnets, it is possible
to redefine the operators Si ± in Eq. (52) by reversing the spin in each second
atom, which assimilates the AFM problem to the FM problem and allows the
consideration of spin waves. However, this procedure creates terms of the type
Si + Sj + and Si − Sj − , where the atoms i and j belong to different sublattices. These
terms go beyond straightforward spin-wave theory, which exclusively involves
Si + Sj − and Si − Sj + . The additional terms yield a quantum-mechanical mixture
of the two sublattices and require an additional diagonalization procedure known as
Bogoliubov transformation [3]. The sublattice admixture reduces both the energy
of the AFM state and the sublattice magnetizations, the latter meaning that the
↑ sublattice acquires some ↓ character and vice versa.
The ground-state energy is discussed most conveniently by starting from two
interacting spins having S = 1/2, as described by Eq. (26). For a given AFM
exchange J < 0, the energies of the FM and AFM states scale as S2 = 1/4 and
–S(S + 1) = −3/4, respectively. More generally, the AFM energy is proportional
to –S(S + δ), where δ describes the intersublattice admixture and 0 < δ < 1. Spin-
wave theory yields δ = 0.363 for the linear chain, δ = 0.158 for the square lattice,
and δ = 0.097 for the simple-cubic lattice. The one-dimensional value is close to
the exact result δ = 0.386 for S = 1/2. A rough estimate for the relative reduction
of the sublattice magnetization in hypercubic magnets (z = 2d) is 0.15/S(d − 1),
corresponding to sublattice magnetizations of 0%, 70%, and 85% in one-, two-, and
three-dimensional magnets having S = 1/2.
The complete magnetization
 collapse
 in one dimension is caused by the involve-
ment of the integral k−1 dk ∼ kd – 1 dk, which exhibits an infrared (small-k)
divergence in one dimension. The same integral is behind Bloch’s law and the
Wagner-Mermin theorem, and in all cases, the divergence indicates that fluctuations
destroy long-range magnetic order in one dimension. However, the underlying
physics is different: The fluctuations considered in Bloch’s law and in the Wagner-
2 Magnetic Exchange Interactions 95

Mermin theorem are of thermodynamic origin, whereas the present ones are
zero-temperature quantum fluctuations. These fluctuations, which are largest for
S = 1/2, are a correlation effect and therefore difficult to treat in density-functional
calculations.
One example is high-temperature superconductivity (HTSC) in La2-x Srx CuO4 ,
which involves 3d9 states in the Cu-O planes of the oxides [9] and where spin-1/2
quantum fluctuations trigger the formation of Cooper pairs. The parent compound
La2 CuO4 is a strongly correlated antiferromagnetic semiconductor, but Sr doping
drives the system toward a phase transition. The denominator in Eq. (40) becomes
small, meaning that spin fluctuations (antiferromagnetic paramagnons) are strongly
enhanced by Sr doping. Furthermore, both spin-charge separation [9] and critical
slowing down cause the spin fluctuations to evolve very sluggishly, so that they can
play the role of phonons in BCS superconductors.

Frustration, Spin Liquids, and Spin Ice

A number of exotic topics in physics are more or less closely related to exchange
interactions. This subsection discusses both classical and quantum-mechanical
implications of frustration, as well as some related micromagnetic questions.

Frustration Ring configurations with antiferromagnetic interactions and odd num-


bers N of atoms offer intriguing physics. Let us start with theclassical exchange
energy between atoms at Ri and Rj , which is equal to −Jij cos φi − φj . Consider
an equilateral triangle (N = 3) with antiferromagnetic nearest-neighbor interactions
(Fig. 14). If the exchange was ferromagnetic, then φ i = 0 (or φ i = const.) would
simultaneously minimize the energy of all bonds and yield a ground-state energy
of −3J . The corresponding antiferromagnetic solution, of energy −3 |J |, does not
exist, because three antiferromagnetic bonds cannot be simultaneously realized in
a triangle (Fig. 14a). This is referred to as magnetic frustration. Fig. 14b shows
that bond angles of 120◦ , rather than 180◦ , may be realized for all spins, and the
corresponding ground-state energy is −1.5 |J |, somewhat lower than the energy
− |J | of (a).

The classical frustration problem of Fig. 14 is elegantly summarized by a


construction known as Frost’s circle. The approach was developed to describe
the hopping of p electrons in cyclic molecules [89] but can also be applied to
s-state electrons such as those in Fig. 7 and to interatomic exchange, because the
involvement of τ ij and Jij is mathematically equivalent. For a ring of N atoms with
nearest-neighbor exchange J , the energy eigenvalues per atom are

En = –2 J cos (2 π n/N ) (62)

where n = 0, ..., N–1, N. These energies can be arranged on a circle, as exemplified


by the example N = 5 and ferromagnetic coupling (J > 0) in Fig. 14(c). The
96 R. Skomski

Fig. 14 Classical frustration in rings of N atoms: (a) frustrated state (N = 3), (b) ground state
(N = 3), and (c) graphical solution (Frost’s cycle) for N = 5

FM ground state has n = 0 and the energy −2J . However, the figure shows
that energy eigenvalues are not necessarily symmetric with respect to changing
the sign of J . For odd values of N, the AFM ground state is double degenerate
and characterized by nearest-neighbor spin angles (1–1/N) 180◦ , as opposed to the
180◦ expected for ideal antiferromagnetism. The incomplete antiparallelity leads to
a ground-state energy 2J cos (π/N), higher than that of an ideal antiferromagnet
(2J ). This analysis shows exchange in antiferromagnets is different from exchange
in ferromagnets, even in the classical limit.
The quantum-mechanical ground state of the AFM spin-1/2 Heisenberg triangle
is obtained by using Eq. (52) to evaluate the matrix elements between the states
|↑↑↓>, |↑↓↑>, and |↓↑↑>. It is sufficient to consider Sz = 1/2, since none of
the terms in Eq. (52) changes the total spin projection Sz . For example, spin
configurations such as |↑↑↑ > (Sz = 3/2) and |↑↑↓ > (Sz = 1/2) do not mix.
Furthermore, Sz = − 1/2 is equivalent to Sz = +1/2 and does not need separate
consideration. For the three states with Sz = 1/2, the Hamiltonian is
⎛ ⎞ ⎛ ⎞
100 111
3
H = J ⎝0 1 0⎠ − J ⎝1 1 1⎠ (63)
2
001 111

The diagonalization of this matrix is trivial and yields one FM eigenstate (1, 1, 1)
of energy – 3J /2 > 0 and two AFM eigenstates of energy 3J /2 < 0, for example,
|Ψ 1 > = (2, −1, −1) and |Ψ 2 > = (0, 1, −1). Explicitly

1
|1 >= √ (2| ↑↑↓ >−| ↑↓↑ >−| ↓↑↑ >) (64a)
6

and
2 Magnetic Exchange Interactions 97

1
|2 > = √ (| ↑↓↑ >−| ↓↑↑ >) (64b)
2

Since the two AFM states are degenerate, |Ψ > = c1 |Ψ 1 > + c2 |Ψ 2 > is also an
eigenstate; complex numbers c1/2 mean a spin component in the y-direction.
Alternatively, Eq. (64b) is a product of the type (|↑↓>− |↓↑>)⊗|↑> and contains a
maximally entangled AFM singlet |↑↓>− |↓↑>. According to Eq. (63), the corners
of the triangle are equivalent, so that the singlet may be placed on any of the three
bonds.
The spin-1/2 Heisenberg square has the spin projections Sz = (0, ±1, ±2). In the
antiferromagnetic ground state (Sz = 0), there are six spin configurations, namely,
the two Néel states |↑↓↑↓> and |↓↑↓↑>, as well as six non-Néel states with pairs
of parallel spins, such as |↑↑↓↓>. Classically, non-Néel states are not expected to
appear in the ground state, but the diagonalization of the corresponding 6 × 6 matrix
yields an AFM ground-state singlet with a strong admixture of non-Néel character,
namely

1 1 1
|>= | ↑↓↑↓ >+ | ↓↑↓↑ >− √ (| ↑↑↓↓ >+ . . . ) (65)
2 2 8

The total weight of the non-Néel configurations in the ground state is 50%. This
ground state is rather complicated but, unlike Eq. (64), not frustrated.

Quantum-spin Liquids The quantum-mechanical behavior of the structures of


Fig. 14 is liquid-like, similar to the Luttinger liquid of Sect. “Antiferromagnetic Spin
Chains”. Two-dimensional magnets are particularly interesting. Figure 15 shows 2D
lattices with the triangular structural elements required for AFM frustration. The
wave functions are complicated but contain AFM singlets, and there are many ways
of arranging these singlets on a lattice (a). Some materials investigated as QSL
materials, such as ZnCu3 (OH)6 Cl2 (herbertsmithite), form Kagome lattices (b).
Triangular and Kagome lattices exhibit similar frustration behaviors, but Kagome
lattices differ by having low-lying excitations [91].

Frustration is also important in spin-ice materials, such as pyrochlore-ordered


Dy2 Ti2 O7 . The pyrochlore structure consists of tetrahedra whose corner atoms are
magnetic. The strong crystal field forces the moments to lie on lines from the corners
to the centers of the tetrahedra, but to fix the spin direction (inward or outward),
one needs an additional criterion known as the “two-in, two-out” rule. There are
many two-in two-out configurations, which creates a spin-ice situation reminiscent
of Fig. 15a.
There are excited spin-ice states where all spins point inward (four in) or outward
(four out), which yields an accumulation of magnetic charge (south poles or north
poles) in the middle of the tetrahedron. Such accumulations are also known as
magnetic monopoles, but this characterization is misleading. Magnetic monopoles
are high-energy elementary particles having B·dA = 0. Their existence cannot
98 R. Skomski

Fig. 15 Frustrated lattices in two dimensions: (a) triangular lattice and (b) Kagome lattice

Fig. 16 Static magnetic field


sources: (a) magnetic dipole
and (b) “magnetic monopole”

be ruled out, but they have never yet been observed in the universe. Solid-state
“monopoles”
 are no monopoles, because they are formed from dipoles and do not
violate B·dA = 0. For illustrative purposes, the magnetic charges near the ends of
a long bar magnet may be regarded monopoles, but these are not real monopoles but
merely the ends of long dipoles. Figure 16 compares a magnetic dipole (a) with a
putative magnetic monopole (b). The configuration (b) may be created in the form of
a magnetic-dipole layer or by some radial magnetization distribution in a magnetic
material. In any case, it requires compensating south-pole charges inside the sphere,
so that B · dA = 0. As a consequence, the magnetic field is actually zero outside
the sphere, so that the finite-length arrows in (b) are without physical basis.
In the context of new materials, it is important to keep in mind that exchange
interactions are described in terms of Hamiltonians. The equation of motion of any
system is more fundamentally governed by the Lagrangian and its time integral, the
action. The difference can be ignored in flat spaces but is important in curved and
periodic spaces, where it corresponds to the Berry phase. The contributions of the
phase are basically of a ‘zero-Hamiltonian’ type and ignored in this chapter.
2 Magnetic Exchange Interactions 99

Acknowledgments This chapter has benefited from help in details by P. Manchanda and R.
Pathak and from discussions with B. Balamurugan, C. Binek, X. Hong, Y. Idzerda, A. Kashyap,
P. S. Kumar, D. Paudyal, T. Schrefl, D. J. Sellmyer, and A. Ullah. The underlying research
in Nebraska has been supported by DOE BES (DE-FG02-04ER46152) NSF EQUATE (OIA-
2044049), the NU Collaborative Initiative, HCC and NCMN.

References
1. Heisenberg, W.: Zur Theorie des Ferromagnetismus. Z. Phys. 49, 619–636 (1928)
2. Pauling, L., Wilson, E.B.: Introduction to Quantum Mechanics. McGraw-Hill, New York
(1935)
3. Jones, W., March, N.H.: Theoretical Solid State Physics I. Wiley & Sons, London (1973)
4. Skomski, R.: Simple Models of Magnetism. University Press, Oxford (2008)
5. Stoner, E.C.: Collective electron ferromagnetism. Proc. Roy. Soc. A. 165, 372–414 (1938)
6. Slater, J.C.: Ferromagnetism in the band theory. Rev. Mod. Phys. 25, 199–210 (1953)
7. Stollhoff, G., Oleś, A.M., Heine, V.: Stoner exchange interaction in transition metals. Phys.
Rev. B. 41, 7028–7041 (1990)
8. Ashcroft, N.W., Mermin, N.D.: Solid State Physics. Saunders, Philadelphia (1976)
9. Fulde, P.: Electron Correlations in Molecules and Solids. Springer, Berlin (1991)
10. Coulson, C.A., Fischer, I.: Notes on the molecular orbital treatment of the hydrogen molecule.
Philos. Mag. 40, 386–393 (1949)
11. Senatore, G., March, N.H.: Recent progress in the field of electron correlation. Rev. Mod. Phys.
66, 445–479 (1994)
12. Avella, A., Manchini, F. (eds.): Strongly Correlated Systems: Theoretical Methods. Springer,
Berlin (2012)
13. Avella, A., Manchini, F. (eds.): Strongly Correlated Systems: Numerical Methods. Springer,
Berlin (2013)
14. Metzner, W., Vollhardt, D.: Correlated lattice fermions in d = ∞ dimensions. Phys. Rev. Lett.
62, 324–327 (1989)
15. V. I. Anisimov, A. I. Poteryaev, M. A. Korotin, A. O. Anokhin, and G Kotliar, First-principles
calculations of the electronic structure and spectra of strongly correlated systems: dynamical
mean-field theory“, J. Phys.: Condens. Matter 9, 7359–7367 (1997)
16. Karbach, M., Müller, G.: Introduction to the Bethe ansatz I. Comput. Phys. 11, 36–44 (1997)
17. Karbach, M., Hu, K., Müller, G.: Introduction to the Bethe ansatz II. Comput. Phys. 12, 565–
573 (1998)
18. Hubbard, J.: Electron correlations in narrow energy bands. Proc. R. Soc. London Ser. A. 276,
238–257 (1963)
19. Gutzwiller, M.C.: Effect of correlation on the ferromagnetism of transition metals. Phys. Rev.
Lett. 10, 159–162 (1963)
20. Şaşıoğlu, E., Friedrich, C., Blügel, S.: Effective Coulomb interaction in transition metals from
constrained random-phase approximation. Phys. Rev. B. 83., 121101R-1-4 (2011)
21. Zaanen, J., Sawatzky, G.A., Allen, J.W.: Band gaps and electronic structure of transition-metal
compounds. Phys. Rev. Lett. 55, 418–421 (1985)
22. Brinkman, W.F., Rice, T.M.: Application of Gutzwiller’s Variational method to the metal-
insulator transition. Phys. Rev. B. 2, 4302–4304 (1970)
23. Hund, F.: Atomtheoretische Deutung des Magnetismus der seltenen Erden. Z. Phys. 33, 855–
859 (1925)
24. Herzberg, G.: Atomic Spectra and Atomic Structure. Dover, New York (1944)
25. Condon, E.U., Shortley, G.H.: The Theory of Atomic Spectra. University Press, Cambridge
(1959)
26. Taylor, K.N.R., Darby, M.I.: Physics of Rare Earth Solids. Chapman and Hall, London
(1972)
100 R. Skomski

27. Goodenough, J.B.: Magnetism and the Chemical Bond. Wiley, New York (1963)
28. Anderson, P.W.: Exchange in insulators: Superexchange, direct exchange, and double
exchange. In: Rado, G.T., Suhl, H. (eds.) Magnetism I, pp. 25–85. Academic Press, New York
(1963)
29. Mattis, D.C.: Theory of Magnetism. Harper and Row, New York (1965)
30. Skomski, R., Zhou, J., Zhang, J., Sellmyer, D.J.: Indirect exchange in dilute magnetic
semiconductors. J. Appl. Phys. 99., 08D504-1-3 (2006)
31. Zener, C.: Interaction between the d-shells in the transition metals. II. Ferromagnetic com-
pounds of manganese with perovskite structure. Phys. Rev. 82, 403–405 (1951)
32. Brueckner, K.A., Sawada, K.: Magnetic susceptibility of an electron gas at high density. Phys.
Rev. 112, 328 (1958)
33. Bloch, F.: Bemerkung zur Elektronentheorie des Ferromagnetismus und der elektrischen
Leitfähigkeit. Z. Physik. 57, 545–555 (1929)
34. Janak, J.F.: Uniform susceptibilities of metallic elements. Phys. Rev. B. 16, 255–262 (1977)
35. Wohlfarth, E.P.: Very weak itinerant Ferromagnets; application to ZrZn2 . J. Appl. Phys. 39,
1061–1066 (1968)
36. Mohn, P.: Magnetism in the solid state. Springer, Berlin (2003)
37. Mohn, P., Wohlfarth, E.P.: The curie temperature of the ferromagnetic transition metals and
their compounds. J. Phys. F. 17, 2421–2430 (1987)
38. Balasubramanian, B., Manchanda, P., Skomski, R., Mukherjee, P., Das, B., George, T.A.,
Hadjipanayis, G.C., Sellmyer, D.J.: Unusual spin correlations in a nanomagnet. Appl. Phys.
Lett. 106., 242401-1-5 (2015)
39. Sommerfeld, A., Bethe, H.A.: Elektronentheorie der Metalle.”, Ferromagnetism. In: Flügge, S.
(ed.) Handbuch der Physik, Vol. 24/II, pp. 334–620. Springer, Berlin (1933)
40. Cardias, R., Szilva, A., Bergman, A., Di Marco, I., Katsnelson, M.I., Lichtenstein, A.I.,
Nordström, L., Klautau, A.B., Eriksson, O., Kvashnin, Y.O.: The Bethe-Slater Curve Revisited;
New Insights from Electronic Structure Theory. Sci. Rep. 7, 4058-1-11 (2017)
41. Skomski, R., Coey, J.M.D.: Permanent magnetism. Institute of Physics, Bristol (1999)
42. Sutton, A.P.: Electronic structure of materials. Oxford University Press (1993)
43. Coey, J.M.D.: New permanent magnets; manganese compounds. J. Phys.: Condens. Matter. 26,
064211-1-6 (2014)
44. Snow, R.J., Bhatkar, H., N’Diaye, A.T., Arenholz, E., Idzerda, Y.U.: Large moments in bcc
Fex Coy Mnz ternary alloy thin films. Appl. Phys. Lett. 112, 072403 (2018)
45. Kashyap, A., Pathak, R., Sellmyer, D.J., Skomski, R.: Theory of Mn-based high-magnetization
alloys. IEEE Trans. Magn. 54, 2102106-1-6 (2018)
46. Wollmann, L., Chadov, S., Kübler, J., Felser, C.: Magnetism in cubic manganese-rich Heusler
compounds. Phys. Rev. B. 90, 214420-1-12 (2014)
47. Skomski, R., Manchanda, P., Kumar, P., Balamurugan, B., Kashyap, A., Sellmyer, D.J.:
Predicting the future of permanent-magnet materials. IEEE Trans. Magn. 49, 3215–3220
(2013)
48. Anderson, P.W.: Concepts in Solids: Lectures on the Theory of Solids. W. A, Benjamin, New
York (1965)
49. Kohn, W.: Nobel lecture: electronic structure of matter—wave functions and density function-
als. Rev. Mod. Phys. 71, 1253–1266 (1999)
50. Perdew, J.P., Burke, K., Ernzerhof, M.: Generalized gradient approximation made simple. Phys.
Rev. Lett. 77, 3865–3868 (1996)
51. Runge, E., Zwicknagl, G.: Electronic structure calculations and strong correlations: a model
study. Ann. Physik. 5, 333–354 (1996)
52. Skomski, R., Manchanda, P., Kashyap, A.: Correlations in rare-earth transition-metal perma-
nent magnets. J. Appl. Phys. 117, 17C740-1-4 (2015)
53. Lieb, E.H.: Density functionals for coulomb systems. Int. J. Quantum Chem. 24, 243–277
(1983)
54. Smart, J.S.: Effective Field Theories of Magnetism. Saunders, Philadelphia (1966)
2 Magnetic Exchange Interactions 101

55. Anisimov, V.I., Aryasetiawan, F., Lichtenstein, A.I.: First-principles calculations of the elec-
tronic structure and spectra of strongly correlated systems: the LDA+ U method. J. Phys.:
Condens. Matter. 9, 767–808 (1997)
56. Liechtenstein, A.I., Katsnelson, M.I., Antropov, V.P., Gubanov, V.A.: Local spin density
functional approach to the theory of exchange interactions in ferromagnetic metals and alloys.
J. Magn. Magn. Mater. 67, 65–74 (1987)
57. Skomski, R., Balamurugan, B., Manchanda, P., Chipara, M., Sellmyer, D.J.: Size dependence
of nanoparticle magnetization. IEEE Trans. Magn. 53, 1–7 (2017)
58. Kondo, J.: Resistance Minimum in Dilute magnetic Alloys. Progr. Theor. Phys. 32, 37–49
(1964)
59. Wijn, H.P.J. (ed.): Magnetic properties of metals: d-elements, alloys, and compounds. Springer,
Berlin (1991)
60. Yeomans, J.M.: Statistical mechanics of phase transitions. University Press, Oxford (1992)
61. Wilson, K.G.: The renormalization group and critical phenomena. Rev. Mod. Phys. 55, 583–
600 (1983)
62. Duc, N.H., Hien, T.D., Givord, D., Franse, J.J.M., de Boer, F.R.: Exchange interactions in rare-
earth transition-metal compounds. J. Magn. Magn. Mater. 124, 305–311 (1993)
63. Kittel, C.: Introduction to Solid-State Physics. Wiley, New York (1986)
64. Kumar, P., Kashyap, A., Balamurugan, B., Shield, J.E., Sellmyer, D.J., Skomski, R.: Permanent
magnetism of intermetallic compounds between light and heavy transition-metal elements.
J. Phys.: Condens. Matter. 26, 064209-1-8 (2014)
65. Coey, J.M.: Magnetism and magnetic materials. University Press, Cambridge (2010)
66. Fischer, K.-H., Hertz, A.J.: Spin glasses. University Press, Cambridge (1991)
67. Moorjani, K., Coey, J.M.D.: Magnetic glasses. Elsevier, Amsterdam (1984)
68. Koehler, W.C.: Magnetic properties of rare-earth metals and alloys. J. Appl. Phys. 36, 1078–
1087 (1965)
69. Stevens, K.W.H.: A note on exchange interactions. Rev. Mod. Phys. 25, 166 (1953)
70. Dzyaloshinsky, I.: A thermodynamic theory of ’weak’ ferromagnetism of antiferromagnetics.
J. Phys. Chem. Solids. 4, 241–255 (1958)
71. Moriya, T.: Anisotropic Superexchange interaction and weak ferromagnetism. Phys. Rev. 120,
91–98 (1960)
72. Bak, P., Jensen, H.H.: Theory of helical magnetic structures and phase transitions in MnSi and
FeGe. J. Phys. C. 13, L881–L885 (1980)
73. Skomski, R.: Nanomagnetics. J. Phys.: Condens. Matter. 15, R841–R896 (2003)
74. R. Skomski, J. Honolka, S. Bornemann, H. Ebert, and A. Enders, “Dzyaloshinski–Moriya
micromagnetics of magnetic surface alloys”, J. Appl. Phys. 105, 07D533-1-3 (2009)
75. Ullah, A., Balamurugan, B., Zhang, W., Valloppilly, S., Li, X.-Z., Pahari, R., Yue,
L.-P., Sokolov, A., Sellmyer, D.J., Skomski, R.: Crystal structure and Dzyaloshinski–Moriya
micromagnetics. IEEE Trans. Magn. 55, 7100305-1-5 (2019)
76. Seki, S., Mochizuki, M.: Skyrmions in Magnetic Materials. Springer International, Cham
(2016)
77. Balasubramanian, B., Manchanda, P., Pahari, R., Chen, Z., Zhang, W., Valloppilly, S.R., Li,
X., Sarella, A., Yue, L., Ullah, A., Dev, P., Muller, D.A., Skomski, R., Hadjipanayis, G.C.,
Sellmyer, D.J.: Chiral Magnetism and High-Temperature Skyrmions in B20-Ordered Co-Si.
Phys. Rev. Lett. 124, 057201-1-6 (2020)
78. Dubowik, J., Gościańska, I.: Micromagnetic approach to exchange Bias. Acta Phys. Pol A.
127, 147–152 (2015)
79. Skomski, R., Kashyap, A., Zhou, J., Sellmyer, D.J.: Anisotropic Exchange. J. Appl. Phys. 97,
10B302-1-3 (2005)
80. de Jongh, L.J., Miedema, A.R.: Experiments on simple magnetic model systems. Adv. Phys.
23, 1–260 (1974)
81. Bander, M., Mills, D.L.: Ferromagnetism of ultrathin films. Phys. Rev. B. 38, 12015–12018
(1988)
102 R. Skomski

82. Ising, E.: Beitrag zur Theorie des Ferromagnetismus. Z. Phys. 31, 253–258 (1925)
83. Binek, C.: Ising-Type Antiferromagnets: Model Systems in Statistical Physics and in the
Magnetism of Exchange Bias. Springer, Berlin (2003)
84. Sachdev, S.: Quantum Phase Transitions. University Press, Cambridge (1999)
85. Skomski, R.: On the Ising character of the quantum-phase transition in LiHoF4 . AIP Adv. 6,
055704-1-5 (2016)
86. Stinchcombe, R.B.: Ising model in a transverse field. I. Basic theory. J. Phys. C. 6, 2459–2483
(1973)
87. Luttinger, J.M.: An exactly soluble model of a many-fermion system. J. of Math. Phys. 4,
1154–1162 (1963)
88. Schofield, A.J.: Non-Fermi liquids. Contemporary Phys. 40, 95–115 (1999)
89. Frost, A.A., Musulin, B.: A mnemonic device for molecular orbital energies. J. Chem. Phys.
21, 572–573 (1953)
90. Balasubramanian, B., Manchanda, P., Pahari, R., Chen, Z., Zhang, W., Valloppilly, S.R., Li,
X., Sarella, A., Yue, L., Ullah, A., Dev, P., Muller, D.A., Skomski, R., Hadjipanayis, G.C.,
Sellmyer, D.J.: Chiral Magnetism and High-Temperature Skyrmions in B20-Ordered Co-Si.
Phys. Rev. Lett. 124, 057201-1-6 (2020)
91. Mendels, P., Bert, F.: Quantum kagome frustrated antiferromagnets: One route to quantum spin
liquids. C. R. Phys. 17, 455–470 (2016)

Ralph Skomski received his PhD from Technische Universität


Dresden in 1991. He worked as a postdoc at Trinity College,
Dublin, and at the Max-Planck-Institute in Halle, before moving
to the University of Nebraska, Lincoln, where he is presently a
Full Research Professor. He is an analytical theorist with primary
research interests in magnetism, nanomaterials, and quantum
mechanics.
Anisotropy and Crystal Field
3
Ralph Skomski, Priyanka Manchanda, and Arti Kashyap

Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
Phenomenology of Anisotropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
Lowest-Order Anisotropies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
Anisotropy and Crystal Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
Tetragonal, Hexagonal, and Trigonal Anisotropies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
Higher-Order Anisotropy Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
Anisotropy Measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
Crystal-Field Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
One-Electron Crystal-Field Splitting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118
Crystal-Field Expansion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
Many-Electron Ions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
Spin-Orbit Coupling and Quenching . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
Rare-Earth Anisotropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
Rare-Earth Ions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
Operator Equivalents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
Single-Ion Anisotropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
Temperature Dependence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
Transition-Metal Anisotropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 145
Perturbation Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 147
Spin-Orbit Matrix Elements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
Crystal Fields and Band Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 150
Itinerant Anisotropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151

R. Skomski ()
University of Nebraska, Lincoln, NE, USA
e-mail: rskomski@neb.rr.com
P. Manchanda
Howard University, Washington, DC, USA
A. Kashyap
IIT Mandi, Mandi, HP, India
e-mail: arti@iitmandi.ac.in

© Springer Nature Switzerland AG 2021 103


J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_3
104 R. Skomski et al.

First-Principle Calculations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154


Case Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156
Other Anisotropy Mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
Magnetostatic Anisotropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
Néel’s Pair-Interaction Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
Two-Ion Anisotropies of Electronic Origin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162
Dzyaloshinski-Moriya Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164
Antiferromagnetic Anisotropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165
Magnetoelastic Anisotropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
Low-Dimensional and Nanoscale Anisotropies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
Surface Anisotropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
Random Anisotropy in Nanoparticles, Amorphous, and Granular Magnets . . . . . . . . . . . . 169
Giant Anisotropy in Low-Dimensional Magnets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171
Appendices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
Appendix A: Spherical Harmonics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
Appendix B: Point Groups . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176
Appendix C: Hydrogen-Like Atomic 3d Wave Functions . . . . . . . . . . . . . . . . . . . . . . . . . . 177
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179

Abstract

Magnetic anisotropy, imposed through crystal-field and magnetostatic inter-


actions, is one of the most iconic, scientifically interesting, and practically
important properties of condensed matter. This article starts with the phe-
nomenology of anisotropy, distinguishing between crystals of cubic, tetragonal,
hexagonal, trigonal, and lower symmetries and between anisotropy contributions
of second and higher orders. The atomic origin of magnetocrystalline anisotropy
is discussed for several classes of materials, ranging from insulating oxides and
rare-earth compounds to iron-series itinerant magnets. A key consideration is the
crystal-field interaction of magnetic atoms, which determines, for example, the
rare-earth single-ion anisotropy of today’s top-performing permanent magnets.
The transmission between crystal field and anisotropy is realized by spin-orbit
coupling. An important crystal-field effect is the suppression of the orbital
moment by the crystal-field, which is known as quenching and has a Janus-head
effect on anisotropy: the crystal field is necessary to create magnetocrystalline
anisotropy, but it also limits the anisotropy in many systems. Finally, we
discuss some other anisotropy mechanisms, such as shape, magnetoelastic, and
exchange anisotropies, and outline how anisotropy is realized in some exemplary
compounds and nanostructures.

Introduction

Magnetic anisotropy means that the energy of a magnetic body depends on the
direction of the magnetization with respect to its shape or crystal axes. It is a quantity
of great importance in technology. For example, it is crucial for a material’s ability to
serve as a soft or hard magnet; governs many aspects of data storage and processing,
3 Anisotropy and Crystal Field 105

such as the areal density in the magnetic recording; and affects the behavior of
microwave and magnetic-cooling materials.
In the simplest case of uniaxial anisotropy, the energy depends on the polar angle
θ but not on the azimuthal angle φ of the magnetization direction:
 
Ea = V K1 sin2 θ + K2 sin4 θ + K3 sin6 θ (1)

Here the Kn are the n-th anisotropy constants and V is the crystal volume. The first
anisotropy constant K1 is often the leading consideration. Ignoring K2 and K3 , the
anisotropy energy is equal to K1 V sin2 θ , and two cases need to be distinguished.
K1 > 0 yields energy minima at θ = 0 and θ = 180◦ , that is, the preferential
magnetization direction is along the z-axis (easy-axis anisotropy). When K1 < 0,
the energy is minimized for θ = 90◦ (easy-plane anisotropy).
The magnitudes of the room-temperature anisotropy constants K1 vary from less
than 5 kJ/m3 in very soft magnets to more than 17 MJ/m3 in SmCo5 . A variety of
rare-earth-free transition-metal alloys have anisotropies between 0.5 and 2.0 MJ/m3 .
YCo5 , where the Y is magnetically inert, has K1 = 5.0 MJ/m3 .
This chapter deals with the phenomenological description and physical origin of
anisotropy. A key question is how magnetic anisotropy depends on crystal structure
and chemical composition. The main contribution to the anisotropy energy of
most materials is magnetocrystalline anisotropy (MCA), which involves spin-orbit
coupling, a relativistic interaction [1]. This mechanism involves two steps. First, the
electrons that carry the magnetic moment interact with the lattice, via electrostatic
crystal field and exchange interactions. Second, the spin-orbit coupling (SOC)
ensures that the spin magnetization actually takes its orientation from the lattice.
In the absence of spin-orbit coupling, anisotropic arrangements of atoms do not
introduce magnetocrystalline
  anisotropy. A good example is the Heisenberg model,
H = –ij J R i –R j S i · S j , which is magnetically isotropic even if the exchange-
bond distribution (Ri – Rj ) is highly anisotropic, for example, in thin films and
nanowires.
Magnetocrystalline anisotropy is not the only contribution. Magnetostatic dipolar
interactions are important in some nanostructured materials and also in materials
where the magnetocrystalline anisotropy is zero by coincidence. Shape anisotropy
(Sect. “Néel’s Pair-Interaction Model”) is a dipole contribution of importance
in some permanent magnets (alnicos), and in magnets with a noncubic crystal
structure, there is also a small dipole contribution to the MCA. The latter is
particularly important in Gd-containing magnets, because Gd3+ ions do not exhibit
anisotropic crystal-field interactions (Sect. “Crystal-Field Theory”) but has a large
dipole moment (S = 7/2).
Magnetic anisotropy is most widely encountered in ferro- and ferrimagnets, but it
is also present in antiferromagnets (Sect. “Magnetoelastic Anisotropy”), disordered
magnets (Sect. “Random Anisotropy in Nanoparticles, Amorphous, and Granular
Magnets”), paramagnets, and diamagnets. An example of an anisotropic diamagnet
is graphite, where the magnitude of the susceptibility is 40 times higher along the
hexagonal c-axis than in the basal plane [2], due to the high mobility of the electrons
in the graphene-like carbon sheets that make up the graphite structure.
106 R. Skomski et al.

Anisotropies in bulk materials and this films are closely related to the orbital
moment L (see below) and therefore to the Bohr-van Leeuwen theorem, which
suggests that a magnetic field acting on electrons does not change the magnetization
of solids. In fact, the field leaves the energy of the electrons unchanged, because the
Lorentz force is perpendicular to the velocity, but nevertheless changes the orbital
moment. Note that the definition and physical interpretation of orbital moments
in solids is a rather recent development, associated with the discoveries of Berry
phase and bulk-boundary correspondence. Berry-phase effects are importantant
in curved and periodic spaces and can be considered as geometrical constant-
energy phenomena. Furthermore, magnetization processes reflect the rotation rather
than creation of atomic moments. Neither the concept of spin roatation nor the
Berry phase where known in the early 20th century, when the theorem was
formulated.
Depending on the relative strength of the spin-orbit coupling compared to the
interatomic interactions (crystal field, exchange, hopping), there are two important
limits. The magnetocrystalline anisotropy of high-performance permanent magnets,
such as SmCo5 and Nd2 Fe14 B, is largely provided by the rare-earth 4f electrons
[3, 4]. These electrons are close to the nucleus (R ≈ 0.5 Å), which means that they
exhibit a strong spin-orbit coupling (of the order of 200 meV) but do not exhibit
much interaction with the crystalline environment (of the order of 10 meV). The
orbits of the 4f electrons, as well as the charge distribution n4f (r), are determined
by Hund’s rules. Specifically, Hund’s first rule, which has its origin in the Pauli
principle, states that the total spin S is maximized. The remaining degeneracy with
respect to total orbital moment L is removed by Hund’s second rule, which means
that the orbital moment L is maximized due to intra-atomic exchange. Finally,
Hund’s third rule describes how the spin-orbit interaction rigidly couples orbital-
moment vector L to the spin vector S.
Due to Hund’s third rule, a change in the magnetization angles θ and φ
yields a rigid rotation of the charge distribution n4f (r). Figure 1 explains how this

Fig. 1 Basic mechanism of magnetocrystalline anisotropy, illustrated by an Sm3+ ion (blue) in a


tetragonal environment (yellow). The Sm spin (arrow) is rigidly coupled to the prolate Hund’s-rules
4f charge cloud of the Sm, and the anisotropy energy is equal to the electrostatic interaction energy
between the Sm3+ electrons and the electrostatic crystal field. Due to electrostatic repulsion, the
energy of (a) is lower than (b), and the anisotropy is easy-axis (K1 > 0). If the prolate Sm3+ ion was
replaced by an oblate Nd3+ ion, the situation would be reversed to easy-plane anisotropy (K1 < 0)
3 Anisotropy and Crystal Field 107

rotation translates into magnetocrystalline anisotropy. The electric field created by


neighboring atoms in the crystal (yellow) is weaker than the spin-orbit coupling and
has little effect on the 4f electronic structure, but the electrostatic interaction of the
4f shell with the crystal causes the energy to depend on θ and φ. Atomic crystal-field
charges are normally negative, which amounts to a repulsive interaction between
4f electrons and neighboring atoms. Figure 1 shows an Sm3+ ion with a prolate
charge distribution. For the tetragonal crystal environment shown in the figure, the
electrostatic energy of (a) is lower than (b), corresponding to easy-axis anisotropy
(K1 > 0).
In the opposite limit of the Fe-series transition-metal magnets, the spin-orbit
interaction is much weaker than the interatomic interactions involving 3d electrons.
The 3d orbits are therefore determined by the crystal field and interatomic hopping,
and the spin-orbit coupling is a small perturbation unless the unperturbed energy
levels are accidentally degenerate. In other words, crystal-field interactions deter-
mine rare-earth anisotropy but are merely an important starting consideration for
the determination of iron-series transition-metal anisotropies.
This chapter starts with the phenomenology of magnetic anisotropy (Sect. “Phe-
nomenology of Anisotropy”), followed by analyses of crystal-field interac-
tions (Sect. “Crystal-Field Theory”), rare-earth anisotropy (Sect. “Rare-Earth
Anisotropy”), and transition-metal anisotropy (Sect. “Other Anisotropy Mecha-
nisms”). The last section deals with some special topics,

Phenomenology of Anisotropy

Magnetocrystalline anisotropy is usually parameterized in terms of anisotropy


constants, as in Eq. (1). The definition of these constants is somewhat arbitrary,
tailored towards experimental and micromagnetic convenience, and there is some
variation in notation. Another approach is to use anisotropy coefficients of order
2n, obtained by expanding the magnetic energy in spherical harmonics (Appendix
A). For example, the anisotropy coefficients κ 2 0 , κ 4 0 , κ 6 0 roughly correspond
to the uniaxial anisotropy constants K1 , K2 , and K3 , respectively. Anisotropy
coefficients are difficult to access by direct magnetic measurements, but they form
an orthonormal system of functions that do not mix crystal-field contributions of
different orders.
Anisotropy energies per atom, typically 1 meV or less, are much smaller than
the energies responsible for moment formation (about 1000 meV). This means that
the spontaneous magnetization Ms = |M| is essentially fixed and that anisotropy
energies can be expressed in terms of the magnetization angles φ and θ . It is
convenient to choose a coordinate frame where

 
M = Ms sin θ cos φ ex + sin θ sin φ ey + cos θ ez (2)

The respective directions x, y, and z with unit vectors ex , ey , and ez often correspond
to the crystallographic a-, b-, and c-axes in crystals of high symmetry.
108 R. Skomski et al.

Lowest-Order Anisotropies

The first- (or second-)order anisotropy constant K1 often provides a good description
of the anisotropy of magnetic substances where higher-order anisotropy constants
are negligible. There are, however, many exceptions to this rule. For example, the
relative magnitudes of the different Kn depend on crystal structure and temperature,
and the φ-dependence of the anisotropy cannot be neglected in many cases. On
the other hand, simplifications arise because many anisotropy constants are zero
by crystal symmetry. Figure 2 shows the corresponding hierarchy: the number of
anisotropy constants decreases in the direction of the arrows.

Fig. 2 Relations between crystal systems; the arrows point in directions of increasing symmetry
and decreasing numbers of anisotropy constants. The arrows also suggest how degeneracies may
arise. For instance, stretching a cubic crystal so that c > a = b creates a tetragonal crystal, whereas
stretching a hexagonal crystal so that a > b creates an orthorhombic crystal
3 Anisotropy and Crystal Field 109

Let us start with the lowest-order anisotropy. Expansion of the magnetic energy
in spherical harmonics shows that there are five second-order terms Y2 m , corre-
sponding to five anisotropy coefficients κ 2 m , namely, κ 2 −2 , κ 2 −1 , κ 2 0 , κ 2 1 , and
κ 2 2 . However, three Euler angles are necessary to fix the anisotropy axes ex , ey ,
and ez relative to the crystal axes, so there remain only two independent anisotropy
constants. Explicitly

Ea (θ, φ)
= K1 sin2 θ + K1  sin θ cos 2φ
2
(3)
V

where K1  is the lowest-order (or second-order) in-plane anisotropy constant; K1


and K1  are generally of comparable magnitude. Equation (3) can be used for any
crystal, but by symmetry, K1  = 0 for trigonal (rhombohedral), hexagonal, and
tetragonal crystals. Only triclinic monoclinic and orthorhombic crystals (red unit
cells in Fig. 2) have K1  = 0. Examples of such low-symmetry compounds are the
monoclinic 3:29 intermetallics [5, 6], such as Nd3 (Fe1-x Tix )29 .
Cubic symmetry is not compatible with these second-order anisotropy contribu-
tions, which means that K1 = 0 and K1  = 0. For example, the replacement of the
tetragonal environment in Fig. 1 by a cubic environment would mean that the x-,
y-, and z-directions are all equivalent, which can only be achieved if K1 = K1  = 0
in Eq. (3). However, there exists a differently defined fourth-order “cubic” K1 c that
reproduces Eq. (3) for small angles θ . To describe cubic anisotropy, one must add
spherical harmonics so that the sum does not violate cubic symmetry. Based on
Table 16, there are only two independent terms that satisfy this condition:

Ea K1 c   K c
2
= 4 x 2 y 2 + y 2 z2 + z2 x 2 + 6 x 2 y 2 z2 (4)
V r r

where x/r = cosα x , y/r = cosα y , and z/r = cosα z are the direction cosines of the
magnetization. From the power-law behavior of r in Eq. (4), we see that K1 c and
K1 c are fourth- and sixth-order anisotropy constants, respectively. K1 c > 0 favors
the alignment of the magnetization along the [001] cube edges, which is called iron-
type anisotropy, while K1 c < 0 corresponds to an alignment along the [98] cube
diagonals and is referred to as nickel-type anisotropy . The subscript “c” is often
omitted, but when uniaxial and cubic anisotropies need to be distinguished, then it
is better to use Ku and/or K1 c to distinguish the respective anisotropy constants.

Anisotropy and Crystal Structure

The number of anisotropy constants rapidly increases with increasing order and
decreasing symmetry . By definition, Ea (−M) = Ea (M), so that we need even-
order spherical harmonics only (gray rows in Table 16). The maximum number of
anisotropy constants is therefore 5 (up to second order), 14 (up to the fourth order),
and 27 (up to the sixth order). Since the anisotropy axes do not necessarily corre-
110 R. Skomski et al.

spond to the crystallographic axes, three of these anisotropy constants effectively


function as Euler angles to fix the orientation of the anisotropy axes. The anisotropy
axes are known for most crystals of interest in magnetism, which reduces the number
of anisotropy constants to 2 (up to second order), 11 (up to the fourth order), and 24
(up to the sixth order). Anisotropy constants of the eighth- and higher-order occur
in itinerant magnets, for example, (Sect. “Transition-Metal Anisotropy”), but they
are usually very small and rarely considered.
The Euler angles must be considered in crystals with low symmetry. Table 1 lists
the crystal systems, point groups, and space groups for some magnetic substances.
Appendix B gives a complete list of all 32 points and 230 space groups. Triclinic
crystals always need three Euler angles to relate the crystallographic a-, b-, and
c-axes to the magnetic x-, y-, and z-axes. In all other noncubic crystals, the c-
axis is parallel to the z-axis, and one needs at most one Euler angle φ o . This
angle corresponds to a rotation of the crystal around the c-axis, and the identity
cos(β – β o ) = cos (β) cos (β o ) + sin (β) sin (β o ) can be used to get rid of
one in-plane anisotropy constant at the expense of introducing a generally unknown
rotation angle.
As a macroscopic property, magnetic anisotropy is determined by the point
group of the crystal. Most magnetic substances with orthorhombic, tetragonal,
rhombohedral (trigonal), or hexagonal structures belong to the cyclic (C) or dihedral
(D) Schönflies groups. The groups Cn have a single n-fold rotation axis (c-axis),
whereas Cnh and Cnv also have one horizontal and n/2 vertical mirror planes,
respectively. The dihedral groups Dn have an n-fold rotation axis and n/2 additional
twofold rotation axes perpendicular to the c-axis. The cubic crystal system contains
tetrahedral (T) and octahedral (O) Schönflies groups.
Most noncubic crystal structures of interest in magnetism belong to the highly
symmetric Schönflies point groups Cnv , Dn , Dnh , and Dnd , which have φ o = 0. This
includes hexagonal (6 mm, 622, 6/mmm, 62m), trigonal (3 m, 32, 3m), tetragonal
(4 mm, 422, 4/mmm, 42m), and orthorhombic (2 mm, 222, 2/mmm) crystals.
Nonzero values of φ o need to be considered in crystals with point groups Cn ,
Cnh , and Sn . This is the case for all triclinic and monoclinic crystals and for some
hexagonal (6, 6/m, 6), trigonal (3, 3), and tetragonal (4, 4/m, 4) point groups. An
example is the monoclinic 3:29 structure [5], which has the point group C2h .
To elaborate on the role of the point groups, it is instructive to compare Cn and
Cnh with Cnv . Figure 3 shows a top view of a fictitious tetragonal crystal. The
fourfold symmetry axis is clearly visible, and since the horizontal mirror plane
is in the plane of the paper, the figure describes both Cn and Cnh . Some of the
nonmagnetic atoms act as “ligands” (red crosses) and create a crystal field that acts
on the rare-earth ions (blue) and establishes local easy axes (dashed line). The local
3 Anisotropy and Crystal Field 111

Table 1 Crystal systems, point groups, space groups, Strukturbericht notation, and prototype
structures for some compounds of interest in magnetism. Not all the examples are ferromagnetic
Crystal system Point group Space group Examples
Monoclinic C2h (2/m) C2/m D015 (AlCl3 ): DyCl3
Monoclinic C2h (2/m) C2/c B26 (tenorite): CuO
Orthorhombic D2h (mmm) Pnma C37 (Co2 Si): Co2 Si; D011 (cementite):
Fe3 C; goethite: α-FeO(OH); orthorhombic
perovskite: SrRuO3
Tetragonal D4h (4/mmm) P4/mmm L10 (CuAu): PtCo, FePd, FePt, MnAl, FeNi
Tetragonal D4h (4/mmm) P42 /mnm Rutile (C4): CrO2 , MnF2 , TiO2 ; Nd2 Fe14 B
Tetragonal D4h (4/mmm) I4/mmm ThMn12 (D2b ): Sm(Fe11 Ti); Al3 Ti (D022 ):
Al3 Dy, GaMn3
Trigonal D3 (32) P32 12 D04 (CrCl3 ): CrCl3 (P31 12)
Trigonal D3d (3m) R3m C19 (α-Sm): Sm, NbS2 ; Th2 Zn17 : Sm2 Co17 ,
Sm2 Fe17 N3
Trigonal D3d (3m) R3c D51 (corundum): α-Fe2 O3
Hexagonal C6v (6 mm) P63 mc B4 (wurtzite): MnSe
Hexagonal D6h (6/mmm) P63 /mmc A3 (hcp): Co, Gd, Dy; B81 (NiAs): MnBi,
FeS; C7 (MoS2 ): TaFe2 ; C14 (MnZn2
hexagonal laves phase): TaFe2 , Fe2 Mo; C36
(MgNi2 hexagonal laves): ScFe2 ; D019
(Ni3 Sn): Co3 Pt*; PbFe12 O19
(magnetoplumbite): BaFe12 O19 , SrFe12 O19 ;
Th2 Ni17 : Y2 Fe17
Hexagonal D6h (6/mmm) P6/mmm D2d (CaCu5 ): SmCo5
Cubic T (23) P21 3 FeSi (B20): MnSi, CoSi, CoGe
Cubic Th (m3) Pa3 Pyrite (C2): FeS2
Cubic Td (43m) F43m C1b (half-Heusler): MnNiSb
Cubic Td (43m) I43m A12 (α-Mn): Mn
Cubic Oh (m3m) Fm3m A1 (fcc): Ni; B1 (NaCl): CoO, NiO, EuO,
US; D03 (AlFe3 ): Fe3 Si; L21 (cubic
Heusler): AlCu2 Mn; D8a (Th6 Mn23 ):
Dy6 Fe23
Cubic Oh (m3m) Im3m A2 (bcc): Fe, Cr
Cubic Oh (m3m) Pm3m B2 (CsCl): NiAl, FeCo, AlCo, B3
(zincblende): CuCl, MnS, GaAs; E21 (cubic
perovskite): BaTiO3 ; L12 (AuCu3 ): Fe3 Pt
Cubic Oh (m3m) Pn3m C3 (cuprite): CuO2
Cubic Oh (m3m) Fd3m C15 (cubic Laves phase): SmFe2 , TbFe2 ,
UFe2 , ZrZn2 ; H11 (spinel): Fe3 O4
Cubic Oh (m3m) Ia3d Fe3 Al2 Si3 O12 (garnet): Y3 Fe5 O12 ,
Gd3 Fe5 O19
112 R. Skomski et al.

Fig. 3 Top view on a unit


cell of a tetragonal crystal
with C4 or C4h symmetry.
Nonmagnetic ligands (red
crosses) create a crystal field
of low symmetry and local
easy axes (dashed lines) that
are unrelated to the crystal
axes (gray lines). For clarity,
the figure shows only some of
the atoms in the unit cell

crystal field may have very low site symmetry, with local easy axes unrelated to the
a- and b-axes (gray). However, since the local easy axes obey the fourfold rotation
symmetry, the sum of all local anisotropy contributions is fourth-order, and there
is no second-order in-plane contribution. The in-plane anisotropy is of the type
cos(4φ – 4φ o ), and in the present example, the angle φ o ≈ 40◦ is equal to the angle
between the dashed local easy axes and the crystal axes. Going from C4(h) to C4v
introduces vertical mirror planes. These two planes ensure that each local easy axis
of angle φ o has a counterpart with –φ o , so that the net anisotropy directions are now
parallel to the crystal axes.
The picture outlined in Fig. 2 and Table 1 focuses on crystallographic point
groups. A more general approach would be to consider magnetic point groups, as
exemplified by the noncubic (tetragonal) electronic structure of magnets having a
cubic crystal structure and a layered antiferromagnetic spin structure. However, the
layers can lie in any of the equivalent cubic lattice planes, so the magnetic anisotropy
remains cubic.

Tetragonal, Hexagonal, and Trigonal Anisotropies

The anisotropy constants belonging to a given point group can be derived by


applying the symmetry elements of the group to the expansion of the magnetic
energy in terms of spherical harmonics. For example, the fourfold rotation symmetry
of tetragonal magnets, Fig. 3, is compatible with cos4φ terms but not with cos2φ
3 Anisotropy and Crystal Field 113

or cos6φ terms. Up to the sixth order, the anisotropy of magnets with a tetragonal
crystal structure is described by

Ea
= K1 sin2 θ + K2 sin4 θ + K2  sin4 θ cos 4φ + K3 sin6 θ + K3  sin6 θ cos 4φ (5)
V

Without further modification, this equation can be used for the space groups C4v ,
D4 , D4h , and D2d . In particular, all tetragonal compounds listed in Table 1 belong
to the highly symmetric point group D4h . The point groups C4 , C4h , and S4 have
a lower symmetry and require consideration of fourth-order angular shifts φ o . The
anisotropy of orthorhombic crystals differs from Eq. (5) by additional second-order
terms, similar to the K2  term in Eq. (3).
The corresponding anisotropy energy expression for trigonal symmetry is

Ea 
V = K1 sin θ + K2 sin θ + K2 sin θ cos θ cos (3 φ)
2 4 3
  (6)
+ K3 sin φ + K3 sin θ cos (6φ) + K3 sin3 θ cos3 θ cos (3
6 6
φ)

Without modification, this equation can be used for the trigonal point groups
C3v , D3 , and D3d as well as for hexagonal crystals, which can be considered as
degenerate trigonal crystals (Fig. 2). Hexagonal point symmetry is ensured by
putting K2  = K3  = 0, and no further modification is necessary for the point groups
C6v , D6 , D6h , and D3h . The lower-symmetry point groups C3 , C3h , and S3 (trigonal)
and C6 , C6h , C3h , and S6 (hexagonal) require the consideration of an angular
shift φ o in each φ-dependent term. Note that the relationship between trigonal,
rhombohedral, and hexagonal crystals is complicated. The term rhombohedral
denotes the translational symmetry (Bravais lattice), whereas the closely related
term trigonal refers to the point symmetry. Some trigonal crystals have hexagonal
rather than rhombohedral translation symmetry. The trigonal space groups whose
names begin with P (for primitive) are hexagonal, whereas those starting with R
are rhombohedral. For example, Table 1 shows that α-Sm and Sm2 Co17 belong to
the space group R3m and are both trigonal and rhombohedral. Translationally, the
difference between hexagonal (P) and rhombohedral (R) is similar to the difference
between primitive (P) and body-centered (I) cubic crystals, the rhombohedral cell
having two extra lattice points.
Equations (5) and (6) also describe cubic crystals, which can be considered as
degenerate tetragonal or trigonal crystals (Fig. 2). Stretching a cubic crystal along
the [001]-axis yields a tetragonal crystal, whereas stretching it along the [111] cube
diagonal yields a rhombohedral crystal. The tetragonal symmetry axis (θ = 0) is
therefore parallel to the cubic [001] direction, and the anisotropy constants obey
K1 = K1 c , K2 = − 7K1 c /8 + K2 c /8, K2  = – K1 c /8 – K2 c /8, K3 = – K2 c /8, and K3 
= K2 c /8. In the trigonal case, θ = 0 refers to the [111] direction, and the √ in lowest-
order anisotropy constants are K1 = − 3K1 c /2, K2 = 7K1 c /12, and K2  = 2K1 c /3.
114 R. Skomski et al.

Higher-Order Anisotropy Effects

Equations (5) and (6) are relatively easy to use in experimental magnetism and
theoretical micromagnetism. However, unlike spherical harmonics, the energy terms
in these equations are nonorthogonal and mix anisotropy contributions of different
orders 2n. For example, uniaxial anisotropy, Eq. (1), has the following presentation
in terms of spherical harmonics:

0   0  
Ea
= κ22 3cos2 θ − 1 + κ84 35 cos4 θ − 30cos2 θ + 3
0  
V (7)
+ κ16
6
231 cos6 θ − 315 cos4 θ − 105 cos2 θ − 5

Comparison of Eqs. (1) and (7) shows that K1 contains not only second-order
(κ 2 0 ) but also fourth-order (κ 4 0 ) and sixth-order (κ 6 0 ) contributions. In more detail,
K1 = – 3κ 2 0 /2 – 5κ 4 0 – 21κ 6 0 /2, K2 = 35κ 4 0 /8 + 189κ 6 0 /8, and K3 = −231κ 6 0 /16.
In many cases, the only important anisotropy contribution is K1 = −3κ 2 0 /2,
but in some cases κ 2 0 = 0 and K1 are dominated by fourth-order terms. An
important example is Nd2 Fe14 B (tetragonal) in a narrow temperature range below
room temperature, where κ 4 0 causes the sign of K1 to change (Fig. 14(d) in
Sect. “Temperature Dependence” and Ref. 7).
Anisotropy contributions of the same order tend to have similar magnitudes,
which is important for understanding experimental data. For example, the two uni-
axial anisotropy constants K1 and K2 provide a consistent fourth-order description
of hexagonal crystals but not of tetragonal crystals, because the non-uniaxial K2 
term in Eq. (5) is also of the fourth order. For second-order uniaxial anisotropies,
see Sect. “Lowest-Order Anisotropies”.
Higher-order anisotropy constants may have drastic effects if K1 ≈ 0 by
coincidence, for example, due to competing sublattice contributions. For example,
uniaxial anisotropy with K1 < 0 and K2 > − K1 /2 yields easy-cone magnetism,
where the negative K1 makes the c-axis an unstable magnetization direction but the
positive K2 prevents the magnetization from reaching the basal plane (a-b-plane). In
this regime, the preferred magnetization direction lies on a cone around the c-axis,
described by the angle θ c = arcsin (|K1 |/2K2 ). The temperature dependences of
K1 and K2 are generally very different; K2 usually negligible at high temperatures.
As a consequence, the preferential magnetization direction may change as a
function of temperature, which is known as a spin-reorientation transition. A similar
film thickness-dependent transition is observed in films where surface and bulk
anisotropy contributions compete.
The ratio K1 /μo Ms has the dimension of a magnetic field, which makes it
possible to compare anisotropies with applied magnetic fields and coercivities. It is
customary to define the corresponding anisotropy field of K1 -only uniaxial magnets
as

2K1
Ha = (8)
μo Ms
3 Anisotropy and Crystal Field 115

Table 2 First- and Substance K1 (MJ/m3 ) K2 (MJ/m3 ) Structure


second-order anisotropy
Fe 0.048 0.015 Cubic
constants at room temperature
[9–11] Ni −0.005 0.005 Cubic
Co 0.53 0 Hexagonal
Fe3 O4 −0.011 0.028 Cubic
Nd2 Fe14 B 4.9 0.65 Tetragonal
Sm2 Fe17 N3 8.6 1.46 Rhombohedral
Sm2 Fe17 C3 7.4 0.74 Rhombohedral
YCo5 5.8 −0.3 Hexagonal
Y2 Co17 4.0 0.3 Hexagonal
Tm2 Co17 1.6 0.2 Hexagonal
Sm2 Fe14 B −12.0 −0.29 Tetragonal

The anisotropy field is defined in a formal way and does not actually exist inside a
magnet; it is equal to the external field that creates a certain effect on the magnet.
Subject to shape anisotropy corrections (Sect. “Magnetostatic Anisotropy”), the
anisotropy field establishes an upper limit to the coercivity Hc . In practice, Hc  Ha ,
which is known as Brown’s paradox. An approximate relation is Hc = α Ha , where
α  1 is the Kronmüller factor [8, 9].
The inclusion of higher-order anisotropies gives rise to different nonequivalent
anisotropy field definitions. For example, using Eq. (1) and comparing the energies
for θ = 0 and θ = 90◦ lead to Ha = 2(K1 + K2 + K3 )/µo Ms . The initial slope of the
perpendicular magnetization curves yields the same Ha , whereas the nucleation field
of uniaxial magnets is not affected by K2 and K3 , so that Eq. (8) remains valid for
uniaxial magnets of arbitrary order. In cubic magnets, the anisotropy fields for iron-
type anisotropy (K1 > 0) are described by Eq. (8), whereas nickel-type anisotropy
(K1 < 0) yields Ha = − 4 K1 /3μo Ms (Table 2).

Anisotropy Measurements

Sucksmith-Thompson method. The experimental determination of magnetic


anisotropy is easiest if single crystals or c-axis-aligned single-crystalline powders or
thin films are available. The Sucksmith-Thompson method uses a magnetic field H
perpendicular to the c-axis and measures the magnetization M in the field direction
[12]. Starting from Eq. (1), ignoring K3 and adding the Zeeman energy yield the
energy density η(M/Ms ):

M2 M4
η = K1 + K2 − μo MH (9)
Ms 2 Ms 4

where M = Ms sinθ . Minimizing the energy, ∂η/∂M = 0, and dividing the result by
μo M yields
116 R. Skomski et al.

H 2 K1 4 K2
= + M2 (10)
M μo Ms 2 μo Ms 4

Plotting H/M as a function of M2 yields K1 and K2 from the intercept and slope of
the straight line, respectively.
Approach to saturation. Samples are often polycrystalline. In the ideal case
of noninteracting grains with second-order uniaxial anisotropy, the corresponding
random-anisotropy problem can be solved explicitly. The approach to saturation
obeys
 
Ha 2
M(H ) = Ms 1− (11)
15H 2

In practice, this method requires the fitting of the three parameters: Ms , the
sought-for Ha = 2 K1 /μo Ms , and a high-field susceptibility that must be used
to ensure that ∂M/∂H = 0 for H = ∞. Note that Eq. (11) does not predict
the sign of K1 , because both easy-axis and easy-plane ensembles yield the same
asymptotic behavior. Note that Eq. (11) is essentially a random-anisotropy relation
(Sect. “Random Anisotropy in Nanoparticles, Amorphous, and Granular Magnets”).
Torque magnetometry. A single-crystalline magnetic sample experiences a
mechanical torque –∂Ea /∂α, where α is a magnetization angle relative to the crystal
axes. The angle α is varied with the help of a rotating magnetic field, and the torque
is monitored as a function of the field direction, for example, by measuring the
twisting angle of a filament to which the sample is attached. The interpretation
of the torque depends on the crystalline orientation of the sample, but if the
torque axis is parallel to a magnetocrystalline symmetry axis, the corresponding
anisotropy constants are readily obtained as Fourier components of the torque
curves [13].
Magnetic circular dichroism. Single-ion anisotropy is closely related to the
orbital moment and approximately proportional to the latter in iron-series transition-
metal magnets (Sect. “Perturbation Theory”). A direct way to probe orbital (and
spin) moments on an atomic scale is X-ray magnetic circular dichroism (XMCD).
Circular dichroism means that circularly polarized photons pass through the sample
and that the absorption is different for left- and right-polarized light [14–16]. This is
because the orbital moment reflects atomic-scale circular currents that interact with
light. Furthermore, due to spin-orbit coupling, the light also interacts with spin, so
that XMCD can also be used to simultaneously measure the spin moment.

Crystal-Field Theory

Electrons in solids occupy states reminiscent of atomic orbitals, even in metals.


This applies, in particular, to the partially filled inner shells of transition-metal
elements, such as the iron-series 3d shells and rare-earth 4f shells. The electrons in
3 Anisotropy and Crystal Field 117

Fig. 4 Angular dependence of 3d wave orbitals: (a) real eigenfunctions and (b) top view on a
mixture of states constructed from m > ∼ exp (ιmφ) with m = ±2. Red and yellow areas in
(a) indicate regions of positive and negative wave functions ψ, respectively, and the darkness
in (b) indicates the electron density ψ*ψ. The wave functions shown in this figure are all
eigenfunctions of the free atoms, but in solids (b), the crystal field, symbolized by ligands (black
dots), favors real wave functions (top), whereas spin-orbit coupling favors complex wave functions
| ± m > (bottom). Details of this “quenching” behavior will be discussed in Sect. “Spin-Orbit
Coupling and Quenching”

the inner shells, which often carry a magnetic moment, interact with the crystalline
environment. The crystal-field (CF) interaction of the Sm3+ ion in Fig. 4 is
one example, but a similar picture is realized in 3d ions, especially in oxides.
Itinerant magnets, such as 3d metals, require additional considerations, because their
electronic structure is largely determined by interatomic hopping (band formation).
Crystal-field theory had its origin in the study of transition-metal complexes
in the last decade of the nineteenth century [17]. An example was the distinc-
tion between violet and green [Co(NH3 )6 ]3+ Cl3 3− , which indicates energy-level
differences of stereochemical origin. The quantitative crystal-field theory dates
back to Bethe [18], who treated the atoms as electrostatic point charges. Since
then, the crystal-field theory has been extended to include quantum mechani-
cal bonding effects in a generalization are known as ligand-field (LF) theory
[19]. As emphasized by Ballhausen [17], the latter is quantitatively superior
to Bethe’s CF theory but leaves the main conclusions of the latter unchanged.
In practice, the terms are often used interchangeably: the atoms surrounding a
magnetic ion are called ligands in both complexes and solids, and the term ligand
field is sometimes used. Physically, both electrostatic and hybridization effects
contribute to the crystal field (ligand field), even in oxides. The focus of this
section is on the traditional electrostatic crystal-field theory, but some interatomic
118 R. Skomski et al.

hybridization effects will be discussed in the context of itinerant anisotropy


(Sect. “Transition-Metal Anisotropy”).

One-Electron Crystal-Field Splitting

The wave functions and charge distributions of the electrons are obtained from the
Schrödinger equation. Hydrogen-like 3d wave functions are listed in Appendix C.
The angular parts of the wave functions follow from the spherical character of the
intra-atomic potential and are the same for Fe-series 3d, Pd-series 4d, and Pt-series
5d electrons. However, the radial parts differ for the three series, and they also
depend on non-hydrogen-like details of the atomic potentials. Figure 4 shows the
angular distribution of the five 3d orbitals ψ μ (r). In a free atom, the five orbitals
are degenerate, but in solids and molecules, they undergo crystal-field interactions
described by the Hamiltonian:

HCF = V (r) n(r) dV (12)

where V (r) is the crystal or ligand-field potential and n(r) = ψ ∗ (r)ψ(r) refers to
the d or f orbital(s) in question.
To understand crystal-field effects, it is necessary to consider the shape of the
orbitals. Atomic wave functions and charge distributions such as those shown in
Figs. 4 and 1, respectively, have characteristic prolate, spherical, or oblate shapes.
The larger the magnitude of the quantum number m = lz , the more oblate or flatter
the orbitals, as we can in Fig. 4. This is because large orbital moments, m = ± 2
in Fig. 4 , correspond to a pronounced circular electron motion in the plane perpen-
dicular to the quantization axis (z-axis). By contrast, the prolate |z2 > orbital, which
has m = 0 has its electron cloud close to the z-axis. In a crystalline environment, the
different orbital shapes correspond to different electrostatic interactions. Crystal-
field charges are negative [20], so that the interaction between the 3d or 4f electronic
charge clouds and those of the surrounding atoms is repulsive. As a consequence,
the prolate |z2 > orbital prefers to point in interstitial directions between the atomic
neighbors, rather than towards them. The opposite is true for the oblate orbitals with
m = ± 2.
The electrostatic repulsion between the 3d electrons and those of the neighboring
atoms removes the degeneracy of the five 3d levels and yields the famous eg -t2g
splitting in an environment with cubic symmetry. Figure 5(a) shows the |z2 > orbitals
in a cubal environment where the central atom is coordinated by 8 neighbors.
The |z2 > orbital points in an electrostatically favorable direction and has a very
low energy. The charge distribution of the |x2 -y2 > orbital also points in directions
away from the neighboring atoms or ligands, and it can be shown that the
|x2 -y2 > and |z2 > have the same energy, forming a so-called eg doublet. The |xy>,
|xz>, |yz> orbitals are equivalent by symmetry and form a t2g triplet. The charge
3 Anisotropy and Crystal Field 119

Fig. 5 A 3d orbital (z2 ) in


some crystalline
environments: (a) cubal, (b)
octahedral, (c) tetrahedral,
and (d) tetragonally distorted
cubal. Note that (a), (b), and
(c) have cubic symmetry,
whereas (d) is tetragonal

distributions of the triplet orbitals are closer to the ligands, so that the triplet energy
is higher than the doublet energy.
The opposite splitting is realized in an octahedral environment, Fig. 5(b), where
the central atoms are coordinated by six ligand atoms. In this environment, the |x2 -
y2 > and |z2 > orbitals point directly towards the neighboring atoms, whereas the |xy>,
|xz>, |yz> orbitals point in interstitial directions. The tetrahedral environment (c)
has no inversion symmetry but is otherwise very similar to the cubal environment.
Basically, the cubic e-t2 crystal-field splitting is reduced by a factor 2, because
there are only four neighbors. Symmetries lower than cubic partially or completely
remove the eg and t2g degeneracies. Figure 5(d) illustrates this for a tetragonally
distorted cubal environment. Compared to (a), the ligands move towards the basal
plane, which lowers the energy of the |z2 > orbital but raises that of the |x2 -
y2 > orbital. As a consequence, these states no longer form a doublet. Similarly,
the |xz> and |yz> orbitals become somewhat more favorable compared to the
|xy > orbital, because their charge distribution has a substantial out-of-plane
component. This splits the t2g triplet, but |xz> and |yz> remain degenerate, because
the x and y directions are equivalent in a tetragonal crystal.
Figure 6 summarizes the eg -t2g splitting and the evolution of the levels due
to a symmetry-breaking tetragonal distortion. It is important to note that half-
filled (and full) 3d shells have spherical charge distributions and do not interact
with anisotropic crystal fields. Equivalently, the CF interaction leaves the center
of gravity of the 3d levels unchanged. This can be used to gain some quantitative
information about the level splitting. For example, the eg -t2g splitting, also known
120 R. Skomski et al.

Fig. 6 Crystal-field splitting of 3delectrons in cubic and tetragonal environments

as 10Dq, consists of an energy shift of +6Dq for the doublet and a –4Dq shift for
the triplet.
Table 3 lists the crystal-field splittings for the most symmetric point groups in
each crystal system and for axial symmetry. The levels are described by Mullikan
symmetry labels, using t and e for triplets and doublets, respectively [24]. Singlets
are denoted by a or b, depending on whether the reference axis is an n-fold rotation
axis (a) or not (b). The subscripts 1, 2, and 3 indicate C2 symmetries around crystal
axes, and primes ( ) and double-primes ( ) refer to horizontal mirror symmetry
and antisymmetry, respectively. The subscript g (German gerade “even”) denotes
inversion symmetry, which exists for the cubal coordination, Fig. 5(a), but not for
the otherwise very similar tetrahedral coordination, Fig. 5(c). However, the inversion
symmetry of the 3d wave functions means that there are no levels with subscript u
(German ungerade “odd”), so that no confusion arises by dropping the subscript
g [25]. In each crystal system, the complexity of the symmetry labels decreases
with decreasing symmetry. For example, the eg -t2g splitting is limited to the highly
symmetric point group Oh : the respective cubic compounds FeS2 (space group
Th ), MnNiSb (space group Td ), and FeSi (space group T) have eg -tg , e-t2 , and
e-t splittings.
3 Anisotropy and Crystal Field 121

Table 3 Crystal-field splittings of 3d electrons. The colors indicate the crystal-field multiplet
structure: one doublet and one triplet (red), one singlet and two doublets (yellow), three singlets
and one doublet (green), and five singlets (blue). The listed point groups are the most symmetric
ones in each crystal system – less symmetric point groups yield modified symbols, such as missing
subscripts g. In linear molecules (point groups D∞h and C∞v ), the multiplets a1 , e1 , and e2 are
also known as  + , , and , respectively
122 R. Skomski et al.

Crystal-Field Expansion

It is convenient to expand the crystal-field potential V (r) into spherical harmonics


Yl m (θ , φ). The corresponding expansion coefficients Al m are known as crystal-
field parameters and play an important role in crystal-field theory and magnetism.
Treating the ligands (i = 1 ... N) as electrostatic point charges [18] located at Ri
yields the crystal-field potential energy:

e
N
qi
V (r) = − (13)
4πεo | Ri − r |
i=1

This sum is easily converted into a sum of spherical harmonics by exploiting the
identity:

1 4 π rl
= Y1 m∗ (Θ, Φ) Y1 m (θ, φ) (14)
|R−r | l (2 l + 1) R l+1 |m|<l

so long as R > r. Strictly speaking, the l-summation extends from zero to infinity,
but the symmetry of n(r) in Eq. (12) means that the only relevant terms are l = 2, 4
(d-orbitals) and l = 2, 4, 6 (f -orbitals).
Inserting Eq. (14) into Eq. (13) and summing over all ligands leads to the
cancellation of Yl m (θ , φ) terms that are incompatible with the symmetry of the
crystal. For example, cubic crystals have
 
V (r) = 20A4 0 x 4 + y 4 + z4 − 3r 4 /5 (15a)

where the dimensionless crystal-field parameter 4πεo R5 A4 0 /qe is equal to −7/16,


7/18, and 7/36 for the octahedral, cubal, and tetrahedral ligands of Fig. 5, respec-
tively. The r4 term in Eq. (15a) is isotropic and not necessary for the description
of magnetic anisotropy, but it ensures that the center of gravity of the energy is
conserved during crystal-field splitting. Since x2 + y2 + z2 = r2 , Eq. (15a) is
equivalent to
 
V (r) = −40A4 0 x 2 y 2 + y 2 z2 + z2 x 2 − r 4 /5 (15b)

and to any linear combination of Eqs. (15a) and (15b). The structure of this equation
mirrors that of Eq. (4) for the anisotropy of cubic magnets. A third version of Eq.
(15) will be discussed in the context of operator equivalents.
Uniaxial crystal fields are described by
   
V (r) = A2 0 3 z2 − r 2 + A4 0 35 z4 − 35 z2r 2 + 3 r 4
(16)
+A6 0 231 z6 − 315 z4 r 2 + 105 z2 r 4 − 5 r 4
3 Anisotropy and Crystal Field 123

Table 4 Crystal-field Compound A2 0 A4 0


parameters for some noncubic K/ao 2 K/ao 4
rare-earth transition-metal
intermetallics [10] R2 Fe14 B 300 −13
R2 Fe17 34 −3
R2 Fe17 N3 −358 −39

From Eq. (14) we see that the small parameter in the ligand-field expansion is
r/R, that is, the ratio of d-shell radius to interatomic distance. For this reason,
A4 0 is typically smaller than A2 0 by a factor of order (r/R)2 , or about one order
of magnitude. Exceptions are, for example, weakly distorted cubic structures.
Another way of interpreting crystal fields is to expand V (r) into a Taylor series
with respect to x, y, and z. The nonzero expansion coefficients are the crystal-field
parameters Al m , where l denotes the l-th spatial derivative of V (r). In particular,
A2 0 ∼ ∂ 2 V (r) /∂z2 or, in terms of the electric field, A2 0 ∼ ∂Ez /∂z. This means
that A2 0 is essentially an electric field gradient .
The point-charge model accurately describes the symmetry of the crystal field
[20] and yields semiquantitatively correct numerical predictions for a variety of
systems. It was originally developed for insulators but also approximates rare-earth
ions in metals where the electrostatic interaction is screened by conduction electrons
[21]. This surprisingly broad applicability has its origin in the superposition princi-
ple of crystal-field interactions, which states that the effects of different ligand atoms
are additive in very good approximation [20]. Experimentally, crystal-field effects
are measured most directly by spectroscopy, for example, optical spectroscopy or
inelastic neutron scattering, but there are also indirect measurements, such as rare-
earth anisotropy measurements (Table 4).

Many-Electron Ions

A fixed number n of inner-shell electrons of an ion is called a configuration, such


as 3dn and 4fn . In practice, the configuration corresponds to the ions’ charge state.
All rare-earth elements form tripositive ions, R3+ , as exemplified by Sm3+ (4f5 )
and Dy3+ (4f9 ). Some form R2+ shells such as europium in EuO or R4+ in mixed-
valence and heavy-fermion compounds such as CeAl3 [22, 23]. Transition-metal
ions show a greater variety, most commonly T2+ , T3+ , and T4+ , where the ionic
charge is determined by chemical considerations. For example, Fe3 O4 contains both
Fe2+ (3d6 ) and Fe3+ (3d5 ) ions to charge-compensate the O2− anions.
The n electrons are distributed over the available 2 × (2 l + 1) one-electron states
and labeled by sz = ±1/2 and lz = −l, ..., l – 1, l. The relationship between these
electrons is largely governed by the Pauli principle, by Hund’s-rules for electron-
electron interactions, and by spin-orbit coupling. The Pauli principle means that
each real-space d or f orbital can accommodate at most one↑ and one↓ electron.
Subject to the Pauli principle, there are several ways to place n electrons onto the 10
124 R. Skomski et al.

one-electron 3d levels, each combination corresponding to a many-electron state.


These can be divided into terms characterized by well-defined total spin S =  i
si and orbital quantum numbers L =  i li (i = 1 ... n), with each term containing
(2 S + 1) (2 L + 1) states. The terms are usually denoted by 2S + 1 L, where 2S + 1
is the spin multiplicity and L is denoted by as S (L = 0), P (L = 1), D (L = 2), F
(L = 3), G (L = 4), H (L = 5), and I (L = 6). More generally, it is common to use
capital letters for ionic properties, and S, P, D, F are analogous to one-electron states
s, p, d, and f.
An example is the 3d2 configuration, realized, for example, in Ti2+ . The first
electron can occupy any of the 2 × 5 states, leaving nine states for the second
electron. This yields 90/2 = 45 permutations, each corresponding to a two-electron
state. The highest L is achieved by placing two electrons in the lz = 2 state, (↑↓,
−, −, −, −). This yields S = 0 and L = 4, that is, a1 G term containing 9 states.
The wave function (↑, ↑, −, −, −) has S = 1 and L = 3 and therefore belongs to
a3 F term, which contains 21 states. The other 3d2 terms are 1 D (5 states) and 3 P (9
states), and 1 S (1 state). Similar term analyses can be made for all configurations
[17, 24, 26] but will not be discussed here, because in magnetism our main interest
is the ground-state term. A trivial case is 3d1 , which corresponds to a single
term 2 D.
As far as symmetry is concerned, the crystal-field splittings of ions are equal to
those of the one-electron states [17]. For example, the octahedral splitting eg -t2g for
a single d electron corresponds to Eg -T2g in D ions. Table 5 shows basic the CF
splittings of many-electron terms in cubic, tetragonal, and trigonal environments.
The subscript-free symmetry labels A (singlet), B (singlet), E (doublet), and T
(triplet) are of the lowest-symmetry type, and the numbers indicate two or more
distinct levels. Note that most point groups have subscripts (1, 2, g, u) that are
important in spectroscopy but not for the explanation of magnetic anisotropy.
Without interactions, the terms of a configuration would be degenerate. In reality,
the degeneracy is removed by the electron-electron interaction:
  
1 ρ (r) ρ r 
U= dV dV  (17)
4πεo | r − r |

Table 5 Basic branching table for crystal-field splittings of many-electron ions. Both ground-
state and excited terms are included, and the table is not restricted to d electrons. For example, the
free-ion triplet of a single p electron (P) remains unaffected by a cubic crystal field but exhibits a
singlet-doublet splitting in tetragonal and trigonal crystals
Term Cubic CF Tetragonal CF Trigonal CF
S A A A
P T A+E A+E
D E+T A + 2B + E A + 2E
F A+2T A + 2B + 2E 3A + 2E
G A+E+2T 3A + 2B + 2E 3A + 3E
H E+3T 3A + 2B + 3E 3A + 4E
3 Anisotropy and Crystal Field 125

where ρ(r) is the electron charge density. The corresponding term splittings are
large, 1.8 eV for Co2+ , and often dominate the behavior of the ion. The term
energies E(L, S) can be calculated in a straightforward way, by applying the lowest-
order perturbation theory to Eq. (17), E(L, S)=< (L,S) | U | (L, S) >[17, 27].
However, the ground-state term is more easily obtained from Hund’s rules. The
first rule states that the total spin S= i si is maximized. In the above 3d2 example,
there are two terms with maximum S, namely, 3 F and 3 P, both having S=1. Hund’s
second rule acts as a tiebreaker, by favoring large L= i li . Since F and P mean L=3
and L=1, respectively, 3 F is the ground-state term of the 3d2 configuration. Table 6
shows some basic properties of 3d ions; 4f ions will be discussed in the context of
rare-earth anisotropy (Sect. Crystal-Field Theory).
Hund’s first rule yields another simplification: in the ground-state term of the
3d5 configuration, there are five ↑ electrons which occupy the five available orbitals,
lz = −2, ... +1, +2. This yields L =  lz = 0, meaning that empty, half-filled, and
completely filled 3d shells are all S-type ions. This principle carries over to magnetic
anisotropy: from Table 5 we see that S states do not undergo crystal-field splitting
but remain in their highly symmetric degenerate A state. The corresponding charge
distribution is spherical, and the ion does not contribute to the magnetocrystalline
anisotropy (except via admixture with a higher excited state). Figure 7 shows the
level splittings of the ground-state terms of the 3d ions in an octahedral crystal field.
Note the half-shell symmetry of the splittings: aside from the sign, there are only
two nontrivial cases, namely, one electron or hole (d1 , d4 , d6 , d9 ) and two electrons
or holes (d2 , d3 , d7 , d8 ).
The crystal-field interaction is normally weaker than the intra-atomic exchange.
However, very strong crystal fields may negate Hund’s rules and cause a transition to
a low-spin state. For example, octahedrally coordinated Fe2+ has the configuration
3d6 , and, according to Fig. 7, a T2g ground state, that is, t2g (↑↑↑↓)-eg (↑↑). The
two ↑ electrons in the eg -doublet experience a competition between electron-
electron interaction, which favors parallel spin alignment, and the CF, which
favors t2g occupancy. In very strong crystal fields, the electronic structure becomes
t2g (↑↑↑↓↓↓)-eg (empty), and the ion loses its magnetic moment. This is an example
of a high-spin low-spin transition. Aside from d6 , the three ions d4 , d5 , and d7

Table 6 Electronic Ion Example Term L S


configurations of 3d ions. The
3d1 Ti3+ , V4+ 2D 2 1/2
listed terms are the
ground-state terms 3d2 Ti2+ , V3+ 3F 3 1
3d3 V2+ , Cr3+ 4F 3 3/2
3d4 Cr2+ , Mn3+ 5D 2 2
3d5 Mn2+ , Fe3+ 6S 0 5/2
3d6 Fe2+ , Co3+ 5D 2 2
3d7 Co2+ , Ni3+ 4F 3 3/2
3d8 Ni2+ , Pd2+ 3F 3 1
3d9 Cu2+ 2D 2 1/2
126 R. Skomski et al.

Fig. 7 Crystal-field splittings


of the ground-state terms of
3d ions in a weak octahedral
crystal field. The energy unit
Dq is one tenth of the eg -t2g
splitting

undergo a high spin low spin in strong octahedral crystal fields, leading to spin
moments of 2 μB (d4 ) and 1 μB (d5 , d7 ).
It is instructive to plot the term energies as a function of the crystal field,
using the eg -t2g splitting 10Dq to quantify the crystal field in an Orgel diagram.
An extension of the Orgel diagram is the Tanabe-Sugano diagram , where both
the crystal field (Dq) and the term energies are divided by the Racah parameter
B [24]. This parameter links Hund’s second rule, namely, the maximization
of L, to the underlying intra-atomic electron-electron interaction and satisfies
E(3 P) – E(3 F) = 15B. The ground-state energy is used as the energy zero, which
helps to visualize transitions. Figure 8 shows a Tanabe-Sugano diagram where the
ground-state term changes from high spin 5 T2g (blue line) to low spin 1 A1g (red
line).
Any splitting ± E of a degenerate state lowers the energy by about E if the
level is only partially occupied. For example, the tetragonal lattice distortion of
Fig. 6 means that the eg doublet splits into a low-lying a1g state and a b1g state
and a single electron in the eg doublet moves to the a1g level. The resulting crystal-
field energy gain competes against the mechanical energy necessary to tetragonally
distort the crystal. However, the former is linear in strain ε, whereas the latter is
quadratic, so that the CF should always create a small distortion. This is known as
the Jahn-Teller effect.
3 Anisotropy and Crystal Field 127

Fig. 8 Tanabe-Sugano diagram for a 3d6 ion [24]. The energy unit Dq is one tenth of the eg -t2g
splitting and B is the Racah parameter [28]. The vertical line indicates a transition from a high-spin
state (blue) to a low-spin state (red)

Spin-Orbit Coupling and Quenching

The interatomic interactions (U ) remove the degeneracy between different terms


and create ions with well-defined L and S. However, L and S do not interact
and can point in any direction. In reality, they are subject to relativistic spin-
orbit coupling, which causes the terms to split into multiplets of well-defined total
angular momentum J, denoted by 2S + 1 LJ . Figure 9 illustrates the origin of spin-
orbit coupling: the orbital motion of the electron (L) creates a magnetic field that
couples to the electron’s own spin (S). This coupling is important for both isotropic
magnetism (moment formation, and exchange) and magnetic anisotropy. The key
role of spin-orbit coupling in the explanation of magnetic anisotropy was first
recognized and exploited by Bloch and Gentile in 1931 [1].
The quasiclassical model of Fig. 9 correctly reproduces the order of magnitude
of the spin-orbit coupling, aside from a factor 1/2 (Thomas correction). The spin-
orbit coupling may be derived directly from the relativistic Dirac wave equation.
The coupling is a fourth-order term in the Pauli expansion of the relativistic energy,
similar to the v4 term in the equation:

v2 1 1
me c 2 1+ 2
= me c 2 + me v 2 − me v 4 (18)
c 2 8
128 R. Skomski et al.

Fig. 9 Spin-orbit coupling in a free ion (schematic). The orbiting spin acts like a current loop and
creates a magnetic field that acts on the spin. The nucleus does not actively participate in the spin-
orbit coupling but merely serves to curve the trajectory of the electron: a circular racetrack would
do equally well

Electromagnetic effects are added by including scalar and vector potentials


[10, 29]. The result of the calculation is the SOC energy [29]:

3  
Hso = 2 2
s · ∇V × k (19)
2me c

This equation shows that the spin-orbit coupling favors a spin direction perpen-
dicular to both potential gradient and direction of motion. For example, electrons
in thin films experience a Rashba effect [30], and there is a small interstitial
contribution to the magnetocrystalline anisotropy [31]. The Rashba effect means
that electrons of wave vector k move in the film plane and experience a potential
gradient perpendicular to the film, which naturally occurs due to broken inversion
symmetry at thin-film surfaces and interfaces. According to Eq. (19), the spin then
prefers to lie in the plane, in one direction perpendicular to k. In the opposite in-
plane spin direction, the energy is enhanced, which is referred to as Rashba splitting
of the electron levels.
The potential gradient is most pronounced near the atomic nuclei, and for
hydrogen-like 1/r potentials

2 Ze2
Hso = l·s (20)
2
2me c 4πεo r 3
2

Using Appendix C, we can evaluate the average <1/r3 > and obtain Hso = ξ l · s.
Here ξ is the spin-orbit coupling constant:

2 Z 4 e2 1
ξ= 2 2 3
  (21)
2me c ao 4πεo n3 l l + 1/ (l + 1)
2
3 Anisotropy and Crystal Field 129

It is instructive to discuss relativistic phenomena in terms of Sommerfeld’s fine-


structure constant, α = e2 /4πεo c ≈ 1/137. Electrons in atoms and solids have
velocities of the order of v = αc, so that from Eq. (18):

v2 1 1
me c 2 1 + = me c 2 + me α 2 c 2 − me α 4 c 4 (22)
c2 2 8

Similarly, Eq. (21) becomes

me 4 4 2 1
ξ= Z α c +   (23)
2 1
n l l + /2 (l + 1)
3

This equation captures the relativistic nature of spin-orbit coupling and magnetic
anisotropy. In terms of powers of α, ξ is a small relativistic correction, similar to
the v4 term in Eqs. (18) and (22), but Z, which is largest for inner-shell electrons in
heavy elements, greatly enhances the effect in partially filled shells. Tables 7 and 8
show values of spin-orbit coupling constants ξ for 3d, 4d, 5d, 4f, and 5f elements,
obtained from Hartree-Fock calculations [32, 33]. A comparison of experimental
data and theoretical predictions indicates that these tables have an accuracy of the
order of 10% [32–34]. Furthermore, ξ somewhat increases with ionicity [34]: going
from T2+ to T3+ and T+ , respectively, changes the SOC constant of late 3d elements
by about ±10%.
There are two limits for many-electron spin-orbit coupling. Russell-Saunders
coupling means that the ion has well-defined values of L =  i li and S =  i si .
They are good quantum numbers, and the SOC is a weak perturbation. This limit
is realized when the intra-atomic interactions are stronger than ξ . In the opposite
limit of j-j coupling, the i-th electron first experiences a one-electron SOC so that
J =  i (li + si ). Most solid-state magnetism involves Russell-Saunders coupling, but
j-j coupling is important in two limits: (a) low-lying levels of very heavy elements,
such actinides, and (b) excited levels of most elements, except very light ones. In (a),
the j-j coupling is imposed by the large λ in heavy atoms, whereas in (b), it reflects
the increased electron separation in excited states. In many cases, Russell-Saunders

Table 7 Spin-orbit coupling ξ (meV) ξ (meV) ξ (meV)


constants for electrons in the
d1 Sc 10 Y 32 La 69
partially filled dipositive 3d,
4d, and 5d transition-metal d2 Ti 15 Zr 48 Hf 196
ions. ξ of the inner 1s, 2s, and d3 V 22 Nb 65 Ta 244
2p electrons in heavy d4 Cr 31 Mo 84 W 302
elements is much stronger d5 Mn 41 Tc 106 Re 360
than the values in this table, d6 Fe 53 Ru 129 Os 419
but closed shells do not
exhibit a net spin-orbit d7 Co 68 Rh 156 Ir 485
coupling d8 Ni 86 Pd 186 Pt 556
d9 Cu 106 Ag 221 Au 633
130 R. Skomski et al.

Table 8 Spin-orbit coupling ξ (meV) ξ (meV)


constants ξ for tripositive 4f
f1 Ce 85 Th 197
and 5f transition-metal ions
[32, 33] f2 Pr 102 Pa 234
f3 Nd 120 U 271
f4 Pm 139 Np 308
f5 Sm 159 Pu 348
f6 Eu 182 Am 390
f7 Gd 205 Cm 432
f8 Tb 230 Bk 478
f9 Dy 257 Cf 525
f10 Ho 286 Es 576
f11 Er 318 Fm 629
f12 Tm 352 Md 686
f13 Yb 422 No –
f14 Lu 143 Lw –

coupling yields the correct multiplet structure but j-j coupling causes quantitative
deviations in the level spacing. For example, the j-j coupling effect on the 3 P2 -3 P0
splitting is negligible in C, about 20% in Si, and dominates in Ge, Sn, and Pb [26].
Magnetic anisotropy reflects low-lying excitations, and Russell-Saunders coupling
therefore applies to both transition metals and rare earths.
The Russell-Saunders coupling establishes the vector model, where J = L + S.
Using the Hund’s-rules ground-state terms to evaluate  i ξ i li · si yields the ionic
spin-orbit coupling Λ L·S, where Λ = ± ξ /2S for less and more than half-filled
shells, respectively [17]. The change of sign at half filling yields Hund’s third rule:
for the early elements in each series, J = L – S, and for the late elements, J = L + S.
Each multiplet has 2 J + 1 Zeeman-like intramultiplet levels, Jz = − J, ...,
(J – 1), J, and the degeneracy of these levels is removed by a magnetic field or by the
crystal field. Due to the g-factor of the electron, a magnetic field couples to (L + 2S)
rather than to (L + S). This makes it necessary to project L + 2S onto J, so that
(L + 2S) · J = g J2 . The Landé g-factor of the ion g = 1 for pure orbital magnetism
(L = J) and g = 2 for pure spin magnetism (L = 0). For arbitrary L and S, less and
more than half-filled shells exhibit g = 1 – S/(J + 1) and g = 1 + S/J, respectively.
The exchange between magnetic ions involves spin only, which mandates the use of
the projection S · J = (g − 1) J2 and yields de Gennes factor G = (g – 1)2 J(J + 1).
This makes it possible to write the exchange interaction as J = G Jo , where Jo is a
J-independent Heisenberg exchange constant.
Spectroscopic and magnetic measurements indicate that Hund’s rules are well
satisfied in rare-earth ions (Sect. “Crystal-Field Theory”), but iron-series transition-
metal ions systematically violate them, especially the third rule. For example,
g ≈ 2 for iron-series atoms in almost all metallic and nonmetallic crystalline
environments. In other words, the magnetic moments of Fe, Co, and Ni originate
3 Anisotropy and Crystal Field 131

Fig. 10 Quenching of the 3d


orbital moment (schematic).
The crystal field creates an
energy landscape that inhibits
the circular orbital motion of
the electron and leads to the
charge density of Fig. 4(b)

nearly exclusively from the spin of the 3d electrons, and the atoms look as if L = 0.
For example, iron has a magnetization of about 2.2 μB , but only about 5% of this
moment is of orbital origin. This effect is known as orbital-moment quenching.
Quenching was first recognized explicitly by van Vleck in 1937 [35].
Figure 10 illustrates the physics behind this effect, namely, the disruption of
the electron’s orbital motion by the crystal field. Mathematically, the difference
between quenched and unquenched wave functions is that between real and complex
spherical harmonics (Appendix A). Consider the two states |x2 –y2 >∼cos(4φ) and
|xy>∼ sin(4φ), which are shown in the top row of Fig. 7. Using lz = −i∂/∂φ
to calculate <lz > = − i ψ* lz ψ dφ yields <lz > = 0; it is completely
quenched. Pictorially, the electron “oscillates” in the valleys between the CF
potential mountains, as indicated by the dashed line in Fig. 10, and these oscillations
yield no net orbital motion. The respective electron densities <ψ|ψ> for |x2 –y2 > and
|xy>, namely, ρ = 1 + cos(8φ) and ρ = 1–2cos(8φ), exhibit complementary minima
and maxima, and the positioning of mountains decide which of the two densities
yields the lower energy.
Rather than asking why the orbital is quenched, we should therefore ask how
an orbital moment arises in a solid. Unquenched orbitals are described by wave
functions of the type exp (±imφ) = cos(mφ) ± i sin(mφ), or |±2>= |x2 –y2 >± |xy>.
These functions describe an uninhibited orbital motion and yield <lz > = ± 2 in units
of . However, the corresponding electron charge cloud is ringlike, ρ = const., so
that the electron occupies both valley and energetically costly hill regions, rather
than being confined to valleys. The competition between spin-orbit coupling (SOC)
and crystal field (CF) decides whether the orbital moment is quenched. In the 4f
case, the SOC is large, and the orbital motion of the electrons remains essentially
unquenched by the CF, as in Fig. 9. The opposite is true for 3d electrons, where the
SOC is a small perturbation to the CF, leading to nearly complete quenching.
132 R. Skomski et al.

Rare-Earth Anisotropy

The magnetocrystalline anisotropy of permanent-magnet materials, such as


Nd2 Fe14 B and SmCo5 , largely originates from the rare-earth sublattice. K1
values are 4.9 MJ/m3 and 17.0 MJ/m3 , respectively. By comparison, bcc iron has
K1 = 0.05 MJ/m3 [10]. The 4f wave functions are nearly unquenched, so that the
magnetocrystalline anisotropy energy is equal to the crystal-field energy, as in Fig. 1.
The basic physical picture of this single-ion anisotropy is clear, but a few questions
remain; it is necessary to determine the shape of the rare-earth 4f shells or ions and
to quantify the relationship between crystal-field interaction and anisotropy. Another
question is the temperature dependence. Anisotropy energies per ion correspond to
very low temperatures, at most a few kelvins, so the observation of anisotropy at
and above room temperature must be explained (Table 9).

Table 9 Anisotropy, magnetization, and Curie temperature of some rare-earth transition-metal


intermetallics [9, 10, 37]
Substance K1 (RT)a μo Ms (RT) Tc Structure
MJ/m3 T K
YCo5 5.2 1.06 987 Hexagonal (CaCu5 )
SmCo5 17.2 1.07 1003 Hexagonal (CaCu5 )
NdCo5 0.7 1.23 910 Hexagonal (CaCu5 )
Y2 Fe14 B 1.06 1.36 571 Tetragonal (Nd2 Fe14 B)
Pr2 Fe14 B 5.6 1.41 565 Tetragonal (Nd2 Fe14 B)
Nd2 Fe14 B 4.9 1.61 585 Tetragonal (Nd2 Fe14 B)
Sm2 Fe14 B −12.0 1.49 618 Tetragonal (Nd2 Fe14 B)
Dy2 Fe14 B 4.5 0.67 593 Tetragonal (Nd2 Fe14 B)
Er2 Fe14 B −0.03 0.95 557 Tetragonal (Nd2 Fe14 B)
Y(Co11 Ti) −0.47 0.93 940 Tetragonal (ThMn12 )
Sm(Fe11 Ti) 4.9 1.14 584 Tetragonal (ThMn12 )
Y(Fe11 Ti) 0.89 1.12 524 Tetragonal (ThMn12 )
Y2 Co17 −0.34 1.25 1167 Hexagonal (Th2 Ni17 )
Nd2 Co17 −1.1 1.39 1150 Rhombohedral (Th2 Zn17 )
Sm2 Co17 3.3 1.20 1190 Rhombohedral (Th2 Zn17 )
Dy2 Co17 −2.6 0.68 1152 Hexagonal (Th2 Ni17 )
Er2 Co17 0.72 0.91 1186 Hexagonal (Th2 Ni17 )
Y2 Fe17 −0.4 0.84 320 Hexagonal (Th2 Ni17 )
Y2 Fe17 N3 −1.1 1.46 694 Hexagonal (Th2 Ni17 )
Sm2 Fe17 −0.8 1.17 389 Rhombohedral (Th2 Zn17 )
Sm2 Fe17 N3 8.9 1.54 749 Rhombohedral (Th2 Zn17 )
TbFe2 0.013 0.84 730 Cubic (laves)
aK c for TbFe2
1
3 Anisotropy and Crystal Field 133

Rare-Earth Ions

Rare-earth atoms tend to form tripositve ions in both metals and insulators. Since
spin-orbit coupling is very strong for inner-shell electrons in heavy elements, the
4f electrons experience a rigid coupling of their spin and orbital moments, with
unquenched orbitals and Hund’s-rules spin-orbit coupling. Magnetic anisotropy
is an intramultiplet effect, involving the 2 J + 1 magnetic quantum states Jz
of the ground-state multiplet. Excited multiplets have relatively high energies,
with the notable exceptions of Eu3+ and Sm3+ [33]. In the former, this energy
is only about 40 meV, but the ground-state moment of Eu2+ is zero, and the
element often adopts a Eu2+ configuration with half-filled shell and zero anisotropy.
Otherwise, the Eu3+ ion shows strong van Vleck susceptibility: the Eu3+ moment
is zero in its J = 0 ground-state multiplet, where the contributions from S = 3
and L=3 cancel, but the first-excited multiplet (7 F1 , J = 1) is only 330 K
above the 7 F0 ground-state multiplet. In the case of Sm3+ , the splitting between
the ground-state multiplet (6 H5/2 ) and the first-excited multiplet (6 H7/2 ) is about
100 meV (∼1000 K) [33], so that interatomic interactions and thermal excitations
yield some admixture of 6 H7/2 character (J-mixing). The focus of this section
is on ground-state multiplets, with a brief discussion of the excited Sm multi-
plet.
To determine the crystal-field energy, it is first necessary to specify the shape of
the 4f shells. Why is the Sm3+ ion in Fig. 1 prolate rather than oblate? Interchanging
oblate and prolate shapes changes the sign of K1 and has far-reaching implications.
A tentative answer is provided by the angular dependence of the (real) one-electron
4f wave functions, which are shown in Fig. 11. States with m = ±3, favored by
Hund’s second rule, are prolate, whereas the m = 0 state is oblate. The strong spin-
orbit coupling then creates axially symmetric superpositions exp.(±mφ) from states
with equal |m|, and Hund’s rules determine how the one-electron orbitals combine
to yield many-electron orbitals.
Like any electric charge distribution, the many-electron 4f shell can be expanded
in spherical harmonics. This multipole expansion provides a successively improved
description of angular features. In the zeroth order, the 4f shell is approximated by
a sphere of charge Q = Qo and does not support any anisotropy. The first-order
corresponds to an electric dipole moment Q1 , which is absent by wave-function
symmetry. The lowest-order electric moment is the second-order quadrupole
moment Q2 , which describes the prolaticity of a charge distribution. Table 10
lists some Hund’s-rules ground-state properties of the tripositive rare-earth ions,
including Q2 .
There is a systematic dependence of the ground-state ionic shape on the number
of 4f electrons. Gd3+ has a half-filled shell and a spherical charge distribution
because Hund’s rules mandate seven ↑ electrons having l = 3, 2, 1, 0, −1, −2,
−3, so that L =  i li = 0 (S-state ion) and Q2 = 0. The other elements follow a
quarter-shell rule: the first and third quarters of the series have oblate ions, and the
second and fourth quarters have prolate ions. This rule is a consequence of particle-
134 R. Skomski et al.

Fig. 11 Angular dependence


of 4f wave functions. Red and
yellow areas indicate regions
of positive and negative wave
functions, respectively. As in
Fig. 4, the wave functions
shown here are the real ones,
and m refers to the wave
functions |m > ∼ exp.(imφ)
from which these wave
functions are constructed

hole symmetry in each half shell: 6 electrons are equivalent to a half-filled shell (7
electrons) with one hole. By Hund’s rules, the first electron(s) in a shell have a large
|m| and are oblate (Fig. 11), corresponding to a negative Q2 . Removing an electron
with a large |m| from a half-filled shell yields one oblate hole, which is the same as
a prolate electron distribution with a positive Q2 .
Table 10 is limited to the quadrupole moment Q2 . Higher-order multipole
moments provide a refined description of the angular dependence of the rare-earth 4f
electron cloud. The third-order octupole and fifth-order triakontadipole moments are
zero by symmetry, but the fourth-order hexadecapole moment (16-pole, Q4 ) and the
sixth-order hexacontatetrapole (64-pole, Q6 ) are generally nonzero. Figure 12 shows
the zoology of the angular dependence of the 4f charge distributions up to the fourth
order. For Hund’s-rules ions, the number of animals is limited by the symmetry of
the wave functions (Figs. 4 and 11), namely, n ≤ 4 for 3d ions and to n ≤ 6 for
4f ions [38]. Furthermore, since the rare-earth 4f electrons are unquenched, the 4f
charge distribution shows axial symmetry, and there are no multipole contributions
Ql m with m = 0. The anisotropy corresponding to the unquenched quadupole
3 Anisotropy and Crystal Field 135

Table 10 Hund’s-rules ground states of 4f ions. The orbitals listed from left to right, lz = 3, 2, 1,
0, −1,–2, −3.

moment of rare-earh ions can be very high, up to a few K per atom in temperature
units [36]. This temperature scale needs to be distinguished from that governing
the temperature dependence of anisotropy constants, which involves interatomic
exchange (Sect. 4.4.4).
136 R. Skomski et al.

Fig. 12 Cartoon illustrating the electrostatic R3+ multipole moments up to the fourth order (Q0 ,
Q2 , and Q4 ). The 4f charge distributions n(r) derive from Figs. 4 and 11 and are both axially and
inversion symmetric

Operator Equivalents

The next step is to quantitatively determine the interaction V (r) n (r) dV
(Sect. “One-Electron Crystal-Field Splitting”) between the crystal field and the
4f charge distribution. This can be done explicitly, in a straightforward but
cumbersome way, but a more elegant method is to use operator equivalents. Both
approaches assume that the crystal field, V (r) or Al m , and the 4f charge distribution,
n(r) or Qn , are known.
The straightforward method is best explained by considering
  the lowest-order
uniaxial limit, where Eq. (16) reduces to V (r) = A2 0 3z2 –r 2 . Substituting this
expression into Eq. (12) yields

  
HCF = A02 3z2 − r 2 n(r)dV (24)
3 Anisotropy and Crystal Field 137

By definition, the integral in this equation is equal to Q2 , so that HCF = A2 0 Q2 .


Equation (24) is exact and easily generalized to other Al m , but the problem remains
to determine Q2 as a function of the ion’s electronic properties and magnetization
angles. For example, the rare-earth crystal field is normally far too weak to affect
the term and multiplet structures, but it usually affects the intramultiplet structure.
These energy values can all be obtained by specifying n(r), but this is a very tedious
method.
A much more elegant approach is the use of Stevens operator equivalents Ol m .
The idea is to replace the real-space coordinates (x/r, y/r, z/r) in expressions such as
Eqs. (24) by the vector operator (Jx , Jy , Jz ), using J± = Jx ± iJy and identities such
as J2 = J(J + 1). The lowest-order noncubic operator equivalents are

O2 0 (J ) = 3 Jz 2 –J (J + 1) (25)

corresponding to 3z2 – r2 and

1  2 
O2 2 (J ) = J+ + J− 2 (26)
2

corresponding to x2 – y2 = ½(x + iy)2 + ½(x – iy)2 . The derivation of higher-order


operator equivalents [33, 38] is straightforward but tedious. For example, the fourth-
order cubic crystal-field expression Eq. (15a) consists of the term
 
1 1
20 x + y + z − 3r /5 = 35z − 30 z r + 3r + 5 (x + iy) + (x–iy)
4 4 4 4 4 2 2 4 4 4
2 2
(27)

and corresponds to O4 0 + 5 O4 4 . Here O4 0 = 35 Jz 4 −  30J (J + 1) Jz +


2

25Jz − 6J (J + 1) + 3J (J + 1) and O4 = 2 J+ + J– . The operators have


2 2 2 4 1 4 4

been tabulated in Refs. 33 and especially 38. It is also possible to define operator
equivalents Ol m (L, Lz ) and related spin Hamiltonians Hspin (S, Sz ) for 3d ions
(Sect. “Transition-Metal Anisotropy”), but the underlying physics is different from
the presently considered rare-earth limit, because L and S are only weakly coupled
(quenching).
The occurrence of Jz and of the ladder operators J± greatly simplifies the
calculation of matrix elements of magnetic ions in a crystal field or exchange field.
For Sm3+ , J = 5/2 yields Jz = ±5/2, ±3/2, and ± 1/2, corresponding to O2 0 = 10,
O2 0 = −2, and O2 0 = −8. The magnitude of the splitting is determined by A2 0 and
by the radial part of n(r), but the evaluation of the Ol m is sufficient to determine the
relative energies, namely, 5:–1:–4 in the present example.
The multipole moments are straightforward linear functions of the operator
equivalents:

Ql = θl < r l >4f Ol 0 (28)


138 R. Skomski et al.

Here the Stevens coefficients θ 2 = α J , θ 4 = β J , and θ 6 = γ J are rare-earth specific


constants that describe how Hund’s rules affect the shape of the R3+ ions [38]. For
example, Sm3+ has α J = 13/32 ·5·7, β J = 2·13/33 ·5·7·11, and γ J = 0. There is no
sixth-order crystal-field interaction for Sm3+ (γ J = 0), because the ground-state
multiplet has J = 5/2 < n/2. However, as mentioned in Sect. “Rare-Earth Ions”,
Sm3+ exhibits a rather unusual low-lying excited multiplet, which has J = 7/2 and
may give a small nonzero γ J contribution due to thermal or quantum mechanical
admixture.
Rare-earth ions in magnetically ordered compounds experience an interatomic
exchange field HJ , so that the rare-earth Hamiltonian becomes [39]

H = l,m Bl m Ol m (J, Jz ) + 2 μo (g–1) J · H J + g μo J · H (29)

Here Bm n = θ n < rn >4f Al m and g J·H describes the comparatively weak Zeeman
interaction and HJ is the exchange field. The quantities L, S, and λ enter this
equation only indirectly, via Hund’s rules and J = L ± S. However, O l m contains
intramultiplet excitations (−J < Jz < J), and the raising and lowering operators
J± in Eq. (26) indicate that off-diagonal crystal fields, such as A2 2 , can change
Jz . To exactly diagonalize Eq. (29), it is necessary to include matrix elements
<Jz | Ol m (J ) | Jz >, where Jz = J’z . These matrix elements are known [38] but
complicate the calculations and the evaluation of the results.
Major simplifications arise if the term involving the exchange energy is much
larger than the CF interaction. This is approximately the case in rare-earth transition-
metal (RE-TM) intermetallics such as Nd2 Fe14 B [39, 40], where the exchange field
is roughly proportional to the RE-TM intersublattice exchange JRT . This strong
exchange field stabilizes states with Jz = ±J, where the sign determines the net
magnetization but does not affect the anisotropy. Intramultiplet excitations, caused
by the operators J± , are effectively suppressed, and only the Ql = θl < r l >4f Ol 0
terms remain to be considered. Furthermore, putting Jz = ±J drastically simplifies
the operator equivalents:

O2 0 = 2 J · (J − 1/2) (30)

O4 0 = 8 J · (J − 1/2) · (J –1) · (J –3/2) (31)

O6 0 = 16 J (J − 1/2) · (J –1) · (J –3/2) · (J − 2) · (J − 5/2) (32)

The corresponding 4f charge distributions are axially symmetric around the


quantization axis (z-axis), and their multiple moments are given by Eq. (28).
Table 11 lists multipole moments derived from Eqs. (30)–(32).
3 Anisotropy and Crystal Field 139

Table 11 Rare-earth Element Q2 /ao 2 Q4 /ao 4 Q6 /ao 4


multipole moments
4f1 Ce3+ −0.748 1.51 0
Ql = θl < r l > Ol 0 for
Jz = J, measured in ml . 4f2 Pr3+ −0.713 −2.12 5.89
ao = 0.529 Å is the Bohr 4f3 Nd3+ −0.258 −1.28 −8.63
radius 4f5 Sm3+ 0.398 0.34 0
4f7 Gd3+ 0 0 0
4f8 Tb3+ −0.548 1.20 −1.28
4f9 Dy3+ −0.521 −1.46 5.64
4f10 Ho3+ −0.199 −1.00 −10.0
4f11 Er3+ 0.190 0.92 8.98
4f12 Tm3+ 0.454 1.14 −4.50
4f13 Yb3+ 0.435 −0.79 0.73

Single-Ion Anisotropy

The anisotropy constants are extracted by rotating the magnetization, that is, by
rotating the 4f charge distribution and calculating the energy. It is convenient to
choose a coordinate frame where J is fixed, that is, to actually rotate the crystal field
around the rare-earth ions. This can be done for each ligand separately, because
crystal fields obey the superposition principle. It starts conveniently from an axial
coordination, R || ez , and the corresponding crystal fields A 2 , A 4 , and A 6 are
referred to as intrinsic crystal fields [20]. In the point-charge model, A2 (R) =
–eq/4πεo R 3 . Due to the axial symmetry of the 4f charge distribution, the rotation
of R into the correct direction relative to the 4f moment involves a polar angle Θ.
For example

1 
A2 0 = A2 3cos2  − 1 (33)
2

describes the rotation of a single ligand. By adding the contributions from all
ligands, one can create any crystal field and any relative orientation between crystal
and magnetic moment. This approach is not limited to uniaxial anisotropy. Equation
(16) is uniaxial, but it contains a z4 term, and by rotating different charges onto
the x- and y-axes, one can create crystal fields of the type x4 + y4 + z4 , which are
cubic. Figure 13 illustrates the rotation of the crystal around the rare-earth ion for a
fourth-order anisotropy contribution. Note that none of the rare-earth ions in Fig. 12
has the ghost shape, but quadrupole moments (Q2 ) do not interact with crystal fields
having fourfold symmetry, so that Fig. 13 actually applies to the UFOs (Ce, Tb) and
to the digesting snakes (Sm, Er, Tb).
Since crystal rotations, for example, Θ = 45◦ in Fig. 13(c), and magnetization
rotations are equivalent, Eq. (33) also describes the energy as a function of the
magnetization angle, that is, the anisotropy energy per rare-earth atom. Explicitly
140 R. Skomski et al.

Fig. 13 Cartoon-like “shaking-ghost” interpretation of fourth-order rare-earth anisotropies. Since


the head, feet, and hands of the ghost are made from negatively charged 4f electrons, electrostatics
favors (a) over (b) and (c). The latter two have the same crystal-field energy, but (c) is easier to
calculate, because it leaves the axis of quantization (arrow) unchanged

1  
Ea = Q2 A2 0 3 cos2 θ − 1 (34)
2

Comparison with Eq. (1) yields

3
K1 = − A2 0 Q2 (35)
2VR

where VR is the crystal volume per rare-earth atom. This equation resolves the rare-
earth anisotropy problem by separating the properties of the 4f shell, described by
Q2 , from the crystal environment, described by A2 0 .
Crystal-field parameters such as A2 0 describe the surroundings of the rare-earth
ion and therefore change little across an isotructural series of compounds with
different rare earths. Examples are A2 0 values of 300 K/ao 2 for R2 Fe14 B, 34 K/ao 2
for R2 Fe17 , and – 358 K/ao 2 for R2 Fe17 N3 . In a given crystalline environment, the
sign of the rare-earth anisotropy depends on whether the ion is prolate or oblate.
A positive K1 is obtained by using oblate ions, such as Nd3+ , on sites where the
crystal-field parameter A2 0 is positive, and prolate ions, such as Sm3+ , in crystalline
environments where A2 0 is negative. This explains the use of neodymium in hard
R2 Fe14 B and RT12 N alloys, whereas samarium is preferred in RCo5 , R2 Fe17 N3 ,
and RT12 intermetallics. The rare-earth ions responsible for the anisotropy must
be magnetic, whereas both magnetic and nonmagnetic ligand atoms contribute to
the crystal field. An interesting example is interstitial nitrogen in Sm2 Fe17 , which
changes the anisotropy from easy-plane to easy-axis [41].
Using volume VR per rare-earth ion as a unit volume, the uniaxial anisotropy
constants are

3 21
K1 = − A2 0 Q2 − 5 A4 0 Q4 − A6 0 Q6 (36)
2 2
3 Anisotropy and Crystal Field 141

35 0 189
K2 = A4 Q4 + A6 0 Q6 (37)
8 8

231 0
K3 = − A6 Q6 (38)
16

Tetragonal magnets also have

1 5
K2 = A4 4 Q4 + A6 4 Q6 (39)
8 8

11
K3 = − A6 4 Q6 (40)
16

whereas hexagonal magnets exhibit only one in-plane term

1
K3 = − A6 6 Q6 (41)
16

Cubic anisotropy can be considered as a special limiting case of tetragonal


anisotropy. Using Eqs. (36)–(40) and dropping terms absent incompatible with cubic
symmetry yields

21 0
K1 c = −5 A4 0 Q4 − A6 Q6 (42)
2

231 0
K2 c = A6 Q6 (43)
2

A striking feature in the last two equations is the absence of independent in-
plane crystal-field parameters, such as A4 4 . While a separate consideration of O4 4 ,
as contrasted to O4 4 ∼ Q4 , is not necessary for rare earths due to the axial symmetry
of the 4f charge clouds, the non-uniaxial CF parameters are not independent but
obey A4 4 = 5A4 0 and A6 4 = − 21A6 0 in cubic symmetry.

Temperature Dependence

Magnetic anisotropy exhibits a temperature dependence that is usually much more


pronounced than that of the spontaneous magnetization. It vanishes at the Curie
point. Figure 14 shows schematic temperature dependences of the anisotropy
constants for some classes of magnetic materials. Anisotropy energies per atom
intrinsically correspond to rather low temperatures, of order 1 K for. Magnetic
anisotropy at or above room temperature therefore requires the help of an inter-
atomic exchange field Hex , which stabilizes the directions of the atomic moments
against thermal fluctuations.
142 R. Skomski et al.

Fig. 14 Temperature dependence of anisotropy (schematic): (a) basic dependence in elemental


magnets, (b) bcc Fe, (c) RCo5 alloys, and (d) Nd2 Fe14 B. The curves in (a) are schematic and less
smooth in practice [70], which reflects subtleties in the electronic structure

Typical rare-earth transition-metal (RE-TM) intermetallics exhibit a strong rare-


earth anisotropy contribution, and for TM-rich intermetallics, this contribution
dominates below and somewhat above room temperature. For example, the low-
temperature anisotropy constants K1 are 26 MJ/m3 for SmCo5 and 6.5 MJ/m3 for
Sm2 Co17 , as compared to room-temperature values of 17 MJ/m3 and 4.2 MJ/m3 .
The exchange field necessary to realize the RE anisotropy contribution is largely
provided by the rare-earth transition-metal (RE-TM) intersublattice exchange JRT ,
rather than the weaker rare-earth rare-earth (RE-RE) exchange [42].
The RE-TM interaction is proportional to J·Hex , that is, the rare-earth ions
behave like paramagnetic ions in an exchange field Hex ∼ JRT MT created by
3 Anisotropy and Crystal Field 143

and proportional to the transition-metal sublattice magnetization MT . Depending


on the sign of Hex , the RE-TM exchange favors Jz = ±J, and the correspond-
ing low-temperature anisotropy is described by Ol m (J, Jz ) =Ol m (J, ±J ), as in
Eqs. 30–32. However, thermal excitation leads to the population of intermedi-
ate intramultiplet levels with |Jz | < J. The randomization becomes important
above some temperature T ∗ ∼ JRT /kB , which is typically of order 100–200 K,
Fig. 14(c–d). Below T*, |Jz | ≈ J, and the anisotropy is only slightly reduced.
Above T*, the rare-earth anisotropy contribution is strongly reduced. In the high-
temperature limit, kB T  JRT , all Jz levels are equally populated and the rare-earth
anisotropy vanishes, because m Ol m (J, m) = 0. The orientations of the 4f charge
clouds are thermally randomized and the net shape of the charge clouds becomes
spherical.
To quantify the temperature dependence, one must evaluate the thermal averages
< Ol m >th . At low temperatures, the quantization of Jz plays a role. The exchange
splitting between Jz = ±J and ± (J – 1) is of order JRT , so that the anisotropy
remains constant or “plateau-like” for T  T*, Fig. 14(c). Above T*, the discrete
level splitting is less important and Jz can be considered as a continuous quantity.
This means that Jz = J cosθ and HRT = –JRT cos (θ ), and the operator equivalents
entering the
 anisotropy  expression simplify to Legendre polynomials, for example,
O2 0 ∼ 12 3cos2 θ − 1 = P2 (cos θ ). The thermal averages

π  
m JRT
< cos θ >= N exp cos θ cosm θ sin θ dθ (44)
kB T
0

are readily evaluated by a high-temperature expansion of the exponential function


and yield the rare-earth anisotropy [43].

JRT 2
K1 (T ) = K1 (0) (45)
15 kB T 2

For anisotropies of arbitrary order m, it can be shown that Km ∼ (JRT /T )2m .


Equation (44) can also be used as a classical estimation for iron-series elements
and for the TM anisotropy contribution in RE-TM intermetallics. However, in
this case, J is not an independent interaction parameter (JRT ) but determined
by the Curie temperature, JTT ≈ kB Tc , and the high-temperature limit of Eq.
(45) is no longer meaningful. For small θ , Eq. (44) leads to <cosm θ > =
1−mk B T /JTT . The exponent m = 1 yields the magnetization, whereas values m>1
are necessary to determine the anisotropy, which is proportional to <Pm > =
1–m (m + 1) kB T /2JTT . These relations correspond to the famous Akulov-Callen
m(m + 1)/2 power laws [44–46]:
 m(m+1)/2
Km/2 (T ) Ms (T )
= (46)
Km/2 (0) Ms (0)
144 R. Skomski et al.

Table 12 First and second-order anisotropy constants at low temperatures (LT) and at room
temperature
Element LT RT Structure Refs.
K2
K1 (MJ/m3 ) (MJ/m3 ) K1 (MJ/m3 ) K2 (MJ/m3 )
Fe 0.052 −0.018 0.048 −0.015 bcc [47]
Co 0.7 0.18 0.41 0.15 fcc [47]
Ni −0.012 0.03 −0.005 −0.002 hcp [47]
Nd2 Fe14 B −18 48 4.3 0.65 tetr. [48]
Pr2 Fe14 B 24 −7 5.6 ≈0 tetr. [48]
Sm2 Fe17 N3 12 3 8.6 1.9 rhomb. [48]

Table 13 Transition-metal Compound K1 K1T K1R Symmetry


and rare-earth contributions
Nd2 Fe14 B 4.9 1.1 3.8 Tetragonal
to the room-temperature
magnetocrystalline anisotropy Sm(Fe11 Ti) 4.8 0.9 3.9 Tetragonal
[10]. All values are in MJ/m3 Sm2 Fe17 N3 8.6 −1.3 9.9 Rhombohedral
Sm2 Co17 3.3 −0.4 3.7 Rhombohedral
SmCo5 17.0 6.5 10.5 Hexagonal

In other words, 2nd-, 4th-, and sixth-order anisotropy contributions are propor-
tional to the third, tenth and 21st powers of the magnetization, respectively. Equation
(46), which is valid up to about 0.65 Tc for Fe, means that higher-order anisotropy
contributions rapidly decrease with increasing temperature. A crude approximation,
based on Ms ∼ (1 – T/Tc )1/3 and used in Fig. 14(a), yields the linear dependence
K1 (T) ≈ K1 (0) (1 – T/Tc ) for the first anisotropy constant K1 of uniaxial magnets
(Table 12).
In summary, the temperature dependence of the anisotropy is a very com-
plex phenomenon. Each crystallographically nonequivalent site generally yields
a different anisotropy contribution with a different temperature dependence, and
the distinction is most pronounced between rare-earth (4f ) and transition-metal
(3d) sites. As a rule of thumb, the RE or TM contributions dominate at low
or high temperatures [40, 49], and their respective temperature dependences are
approximately given by Eqs. (44 and 45) and Eq. (46). In the latter case, K1 ∼ Ms 3
(uniaxial magnets) and K1 ∼ Ms 10 (cubic magnets). Actinide (5f ) anisotropy is
limited by the interatomic exchange, although the spin-orbit coupling is very large,
and its temperature dependence follows that of the magnetization, K1 ∼ Ms [50].
The anisotropy of 3d–5d (and 3d–4d) intermetallics, such as tetragonal PtCo, largely
originates from the heavy transition-metal atoms, but this anisotropy is realized via
spin polarization by the 3d sublattice, roughly corresponding to K1 ∼ Ms 2 [51, 52].
The same dependence is obtained for the two-ion (magnetostatic) contribution to the
magnetocrystalline anisotropy, Sect. “Two-Ion Anisotropies of Electronic Origin”,
because the magnetostatic energy scales as Ms 2 (Table 13).
3 Anisotropy and Crystal Field 145

Transition-Metal Anisotropy

Typical second- and fourth-order iron-series transition-metal anisotropies are


1 MJ/m3 and 0.01 MJ/m3 , respectively, with large variations across individual alloys
and oxides (Tables 14 and 15). The anisotropy constants are often quoted in meV
or μeV per atom, especially in the computational literature dealing with metallic
magnets. A rule-of-thumb conversion for dense-packed iron-series transition-metal
magnets is 1 meV = 14.4 MJ/m3 . In alloys, the anisotropy must be multiplied by
the volume fraction f of the transition metals. For example, the transition-metal
contribution to the anisotropy of transition-metal-rich rare-earth intermetallics
corresponds to f ≈ 0.7, because about 30% of the crystal volume is occupied
by the rare-earth atoms.
The magnetic anisotropy 3d magnets is largely dominated by the degree of
quenching (Sect. “Spin-Orbit Coupling and Quenching”). For oxides, the degree
of quenching was implicitly considered by Bloch and Gentile [1], whereas Brooks
(1940) explicitly considered quenching in itinerant iron-series magnets [53]. An
explanation of quenching in itinerant magnets is provided by the model Hamilto-
nian:

Table 14 Anisotropy, magnetization, and Curie temperature of some oxides [9–11, 37, 63]
K1 μo Ms
Substance (RT) (RT) Tc Structure
MJ/m3 T K
α-Fe2 O3 −0.007 0.003 960 Rhombohedral (Al2 O3 )
γ-Fe2 O3 −0.0046 0.47 863 Cubic (disordered spinel)
Fe3 O4 −0.011 0.60 858 Cubic (ferrite)
MnFe2 O4 −0.003 0.52 573 Cubic (ferrite)
CoFe2 O4 0.270 0.50 793 Cubic (ferrite)
NiFe2 O4 −0.007 0.34 858 Cubic (ferrite)
CuFe2 O4 −0.0060 0.17 728 Cubic (ferrite)
MgFe2 O4 −0.0039 0.14 713 Cubic (ferrite)
BaFe12 O19 0.330 0.48 723 Hexagonal (M ferrite)
SrFe12 O19 0.35 0.46 733 Hexagonal (M ferrite)
PbFe12 O19 0.22 0.40 724 Hexagonal (M ferrite)
BaZnFe17 O27 0.021 0.48 703 Hexagonal (W ferrite)
Y3 Fe5 O12 −0.0007 0.16 560 Cubic (garnet)
Sm3 Fe5 O12 −0.0025 0.17 578 Cubic (garnet)
Dy3 Fe5 O12 −0.0005 0.05 563 Cubic (garnet)
CrO2 0.025 0.56 390 Tetragonal (rutile)
NiMnO3 −0.26 0.13 437 Hexagonal (FeTiO3 )
(La0.7 Sr0.3 )MnO3 −0.002 0.55 370 Rhombohedral (perovskite)
Sr2 FeMoO6 0.028 0.25 425 Orthorhombic
146 R. Skomski et al.

Table 15 Anisotropy, magnetization, and Curie temperature of some transition-metal structures.


PT indicates a structural change near or below the Curie temperature
Substance K1 (RT) μo Ms (RT) Tc Structure Refs.
MJ/m3 T K
Fe 0.048 2.15 1043 Cubic (bcc) [68]
Co (α) 0.53 1.76 1360 Hexagonal [68]
(hcp)
Co (β) −0.05 1.8 1388 Cubic (fcc) [69]
Ni −0.0048 0.62 631 Cubic (fcc) [68]
Fe0.96 C0.04 −0.2 2.0 (PT) Tetragonal [70]
Fe4 N −0.029 1.8 767 Cubic [37]
(modified fcc)
Fe16 N2 1.6 2.7 (PT) Tetragonal [71]
Fe3 B −0.32 1.61 791 Tetragonal [72]
Fe23 B6 0.01 1.70 698 Cubic [73]
(C6 Cr23 )
Fe0.65 Co0.35 0.018 2.43 1210 Cubic (bcc) [37]
FeNi 1.3 1.60 (PT) Tetragonal [68]
(L10 )
Fe0.20 Ni0.80 −0.002 1.02 843 Cubic (fcc) [37]
FePd 1.8 1.37 760 Tetragonal [74]
(L10 )
FePt 6.6 1.43 750 Tetragonal [74]
(L10 )
CoPt 4.9 1.00 840 Tetragonal [74]
(L10 )
Co3 Pta 2.1 1.38 1000 Hexagonal [75]
MnAl 1.7 0.62 650 Tetragonal [74]
(L10 )
MnBi 1.2 0.78 630 Hexagonal [9]
(NiAs)
Mn2 Ga 2.35 0.59 (PT) Tetragonal [76]
(D022 )
Mn3 Ga 1.0 0.23 (PT) Tetragonal [76]
(D022 )
Mn3 Ge 0.91 0.09 (PT) Tetragonal [76]
(D022 )
NiMnSb −6.3 1.10 698 Cubic [37]
(half-Heusler)
Fe7 S8 0.320 0.19 598 Monoclinic [37]
a Extrapolation to fully ordered Co3 Pt has been suggested to yield 3.1 MJ/m3 [67]

   
E1 (k) 0 0 i
H= +λ (47)
0 E2 (k) −i 0

where E1 (k) and E2 (k) are two 3d subbands connected by a spin-orbit matrix
elements ±iλ. The spin-orbit term favors a nonzero net orbital moment, as required
3 Anisotropy and Crystal Field 147

for magnetic anisotropy, but λ ≈ 50 meV is usually much smaller than |E1 (k) –
E2 (k)|, the latter being comparable to the bandwidth W of several eV. However,
even for |E1 (k) – E2 (k)| = W, perturbation theory leads to a small orbital moment
and some residual anisotropy. Furthermore, accidental degeneracies E1 (k) = E2 (k)
yield the eigenvalues ±λ and completely unquenched orbitals. The corresponding
anisotropy energy, about 50 meV per atom, is then huge compared to typical iron-
series anisotropies of 0.1 meV, or about 1 MJ/m3 .
The practical challenge is to add the spin-orbit couplings of all atoms (index i):

Hso = i λi l i · s i (48)

to the isotropic Hamiltonian and to determine the anisotropy contributions from all
bands and k-vectors. To quantitatively determine the anisotropy, this procedure must
be performed for different spin direction s, s || ez and s || ex .

Perturbation Theory

The simplest approach to 3d anisotropy is the perturbation theory as originally


developed by Bloch and Gentile [1] and later popularized by van Vleck [35] and
Bruno [54]. The idea is to consider the Hamiltonian H = Ho + Hso , where Ho (l i )
is the nonrelativistic isotropic part and to consider Hso as a small perturbation. In
the independent-electron approximation, the lowest-order correction proportional to
ξ i = λ is obtained by using the perturbed wave functions |μ k σ >, where μ is a 3d
subband index and the index σ = {↑, ↓} labels the spin direction. Lowest-order
perturbation theory, linear in λ, uses completely quenched orbitals, <li > = 0, and
therefore <li ·si > = < li > ·si = 0.
The next term is quadratic in λ. For a single electron of wave function |μ k σ >,
the corresponding anisotropy energy is


<μk σ |l · s|μ k  σ  ><μ k  σ  |l · s|μkσ >
Ek = λ2 (49)
μ,σ  k Eμ k σ  −Eμkσ

The total second-order anisotropy energy is obtained by summation over all


electrons. Since the SOC leaves the centers of gravity of the one-electron energies
unchanged, there is no net anisotropy contribution from level pairs |μ k σ > and |μ
k  σ  > when both levels are occupied (o) or unoccupied (u). The summation (or
integration) is therefore limited to |μkσ >= |o>and |μ k  σ  > = |u>:
 <o | l · s | u> <u | l · s | o>
E = −λ (50)
o,u Eu −Eo

The numerical determination of the anisotropy constants requires the evaluation


of E for several spin directions s = ½(σ x ex + σ y ey + σ z ez ), where σ x , σ y ,
148 R. Skomski et al.

and σ z are Pauli’s spin matrices. Equation (50) is sometimes reformulated in form
of a statement that the anisotropy energy is proportional to the quantum average
of angular orbital moment. However, this equivalence is limited to small orbital
moments [55] – rare-earth orbital moments are fixed by Hund’s rules (Fig. 13) and
do not change as a function of magnetization direction.
The spin summation is greatly simplified by the factorization of the unperturbed
wave functions, |μ k σ >= |μ k>|σ >, but the k-space summations can only
be performed numerically for most systems. The factorization into |μk> and
|σ > makes it possible to formally perform a summation over |μ k>, |μ k >, and
|σ  > only, leaving the spin s unaveraged. This leads to a spin Hamiltonian of the
general many-electron type:

Hspin = −λ2 S · K · S (51)

where K is a 3 × 3 real-space anisotropy matrix [56, 57]. For uniaxial anisotropy,


Eq. (51) reduces to the anisotropy term:
 
Hspin = D Sz 2 –S (S + 1) /3 (52)

This expression, which mirrors other second-order anisotropy expressions, is


not restricted to magnetocrystalline anisotropy but can also be used for dipolar
anisotropy (see Sect. “Magnetostatic Anisotropy”). It is most useful for 3d ions,
where D is often considered an adjustable parameter. As a rough approximation, Eq.
(52) can also be used for metallic Fe and Co (S ≈ 1). It cannot be used to describe the
anisotropy of independent conduction electrons (S = ½) nor for Ni (S ≈ ½), because
S = ½ yields Sz 2 – S(S + 1)/3 = 0 for Sz = ±½. It is, however, possible to consider
classical averages over a number of electrons, which yields Hspin = D cos2 θ –1/3
and K1 = −D.
Generalizing the perturbation expansion to arbitrary orders n yields anisotropy
constants of the order:

λn
Kn/2 ∼ (53)
Vo (Eo − Eu )n−1

where Vo is the crystal volume per transition-metal atom. This important relation,
known as spin-orbit scaling, was first deduced for lowest-order cubic anisotropy,
where n = 4 and K1 c ∼ K2 ) [1]. In this case, the anisotropy constant scales as
λ4 /A3 , where A is the energy-level splitting in the absence of spin-orbit coupling
(crystal-field splitting or bandwidth). This scaling behavior explains the low cubic
anisotropy of bcc iron (0.05 MJ/m3 ) and Ni (−0.005 MJ/m3 ), as compared to that
of hexagonal Co (0.5 MJ/m3 ) and YCo5 (5 MJ/m3 ).
Equation (53) provides a semiquantitative understanding of transition-metal
anisotropies. In metallic systems, Eu – Eo ∼ W, where the bandwidth W is about
5 eV for iron-series (3d) magnets and somewhat larger for palladium -series (4d),
3 Anisotropy and Crystal Field 149

platinum-series (5d), and actinide (5f ) magnets. The spin-orbit coupling rapidly
increases as the atoms get heavier (Tables 7 and 8), so that heavy transition-
metal elements are able to support very high anisotropies so long as the induced
magnetic moments on the heavy atoms are appreciable. In particular, FePt magnets
are important in magnetic recording [58], but both the low Curie temperature and
the low intrinsic magnetic moment per heavy transition-metal atom make it very
difficult to exploit the high anisotropy of very heavy atoms, up to 1000 MJ/m3 for
actinide compounds such as uranium sulfide [59].
As outlined in Eqs. (49 and 50), quantitative anisotropy calculations require a
summation of all occupied and unoccupied states. This summation involves matrix
elements <o|l·s|u>, which couple wave functions of equal |Lz |, namely, Lz = ±1 and
Lz = ±2, where the quantization axis (z-axis) is parallel to the spin direction (see
below). These matrix elements affect the sign and magnitude of the anisotropy but
do not change its order of magnitude, because they are of order unity. The order of
magnitude of the anisotropy is given by the spin-orbit coupling, which is essentially
fixed for a given element (Tables 7 and 8) and by the denominator Eo – Eu , which
requires a detailed discussion.

Spin-Orbit Matrix Elements

In Eq. (50), the itinerant wave functions |o > and |u > are of the Bloch type and can
therefore be expanded into atomic wave functions. Including spin, there are 10 3d
orbitals per atom, which yield 100 matrix elements <l·s > for each spin direction.
However, the number of independent matrix elements is drastically reduced by
symmetry. First, for the highly symmetric point groups Cnv , Dn , Dnh , and Dnd
(Sect. “Anisotropy and Crystal Structure”), only three spin and orbital-moment
directions need to be considered, namely, x, y, and z. Second, the matrix elements
between ↑↑ and ↓↓ pairs are the same, whereas those for ↑↓ and ↓↑ are equal
and opposite in sign. Third, many of the remaining matrix elements are zero by
symmetry [60].
Explicit matrix elements are obtained by applying equations such as
lˆ z = i(y∂/∂x – x∂/∂y) or lˆ z = −i∂/∂φ to the real or quenched 3d wave functions
of Fig. 4. For example, |xy>∼ sin(4φ) and |x2 –y2 >∼ cos(4φ) yield:

<xy | l̂z |x 2 –y 2 >=2i (54)

This matrix element is imaginary and creates an imaginary (unquenched)


admixture to the wave function, as required for magnetocrystalline anisotropy. For
degenerate |xy> and |x2 –y2 > levels, this matrix element yields the eigenfunctions
exp.(±2iφ) = cos(2φ) ± i sin(2φ), the orbital momentum <lz > = ±2, and the
orbital moment ±2μB .
In terms of Fig. 10, the spin-orbit coupling acts as a perturbation that promotes
hopping from one valley into the next and thereby creates a small net orbital motion.
As outlined above (Sect. “Spin-Orbit Coupling and Quenching”), this motion is
150 R. Skomski et al.

2 2 ˆ
Fig. 15 The three “canonical” d electron orbital-momentum √ matrix elements: (a) <x –y |l z
|xy > = 2i, (b) <xz|lˆ z |yz>= i, and (c) <z2 | l̂x | yz>= 3 i. The dashed lines are out of
the paper plane and visualize the direction of l̂ , but the actual length of the lines is zero, because
all orbitals belong to the same atom

responsible for the small orbital contribution to the magnetic moment of itinerant
magnets, such as Fe, and for the corresponding magnetic anisotropy.
The five 3d orbitals yield a fairly large number of matrix elements such as
that in Eq. (54), but due to symmetry, many of them are zero, and only three
are nonequivalent. Figure 15 illustrates these three “canonical” matrix elements.
Figure 15(a) corresponds to Eq. (54) and is encountered only once, aside from the
conjugate complex value –2i created by interchanging xy and x2 –y2 . The matrix
element of Fig. 15(b) occurs five times, namely, in form of <xz|lˆ z |yz>, <xy|lˆ x |xz>,
<xy|lˆ y |yz>, <x2 –y2 |lˆ x |yz>, and <x2 –y2 |lˆ y |xz>, whereas that of Fig. 15(c) has two
realizations, namely, <z2 |lˆ x |yz> and <z2 |lˆ y |xz>. A physical interpretation of matrix
elements <ψ 1 |l̂ |ψ 2 > is that <ψ 1 |ψ 2 > = 0 but the angular momentum operator
rotates ψ 2 and thereby creates overlap with ψ 1 . The rotation angle is equal to π/m,
where m is the magnetic quantum number of the orbitals, so that π/4 in Fig. 15(a, c)
and π/2 in Fig. 15(b).

Crystal Fields and Band Structure

An important question is the relation between electrostatic crystal-field interaction


and the interatomic hopping that leads to band formation. In the Mott insulator limit
of negligible interatomic hopping, the energy differences Eo – Eu correspond to
the ionic CF level splittings outlined in Sect. “Crystal-Field Theory”. However,
many oxides are Bloch-Wilson insulators, whose insulating character is a band-
filling effect. This means that band effects are not negligible in many or most
oxides. Hybridization-type ligand fields, which include band formation, do not alter
the qualitative physics of crystal-field theory [17] but are often stronger than the
electrostatic crystal fields and strongly affect quantitative anisotropy predictions.
3 Anisotropy and Crystal Field 151

For example, the eg -t2g crystal-field splitting in transition-metal monoxides is of the


order of 1 eV, as compared to 3d bandwidths of about 3 eV [62].
It is important to note that properly set up band structure calculations, from first-
principle (Sect. “First-Principle Calculations”) or based on tight-binding approxi-
mations, automatically include crystal-field effects. This is easily seen by consider-
ing a tight-binding model that is nonperturbative as regards spin-orbit coupling. The
Hamiltonian is

2 2    
H=− ∇ + j Vo r − Rj + j Hso Rj (55)
2me

where the matrix elements of Hso are those of Fig. 15. The lattice periodicity is
accounted for by the ansatz
   
ψkμ (r) = N exp ik · Rj φμ r − Rj (56)
j

where the index μ labels the orbitals, such as |xy↑>. Putting Eq. (56) into Eq. (55)
yields, in matrix notation

Eμμ (k) = Eo δμμ + Aμ δμμ + m exp (ik · Rm ) tμμ (Rm ) + Eso,μμ (57)

Here Eo is the on-site energy, Aμ is the subband-specific crystal-field energy, and


tμμ (Rm ) is the matrix containing the interatomic hopping integrals. The crystal-field
term is easily derived by splitting the potential energy  j Vo (r – Rj ) into an on-site
term Vo (r – Ri ), which enters Eo , and a crystal-field term  j = i Vo (r – Rj ).

Itinerant Anisotropy

Figure 16 shows an explicit example, namely, a monatomic tight-binding spin chain


with two partially occupied ↓ subbands near the Fermi level, namely, |xy > and
|x2 –y2 >, whereas Fig. 17 illustrates the corresponding band structure and anisotropy.
In terms of the fundamental Slater-Koster hopping integrals [64], txy, xy = Vddπ and
tx2–y2 ,x2–y2 = ¾Vddσ + ¼Vddδ , whereas txy ,x2–y2 is zero by symmetry. The ratio
Vddσ :Vddπ :Vddδ is about +6:-4:+1 [65], so that the model creates two cos(ka) bands
of nearly equal widths W ≈ 2Vddπ but opposite slope. The two bands, shown as
dashed curves in Fig. 17, exhibit a crossing at k = π/a.
The solid curves in Fig. 17 differ from the dashed ones by including crystal-field
and spin-orbit interactions. First, the charge distributions of the |x2 –y2 > orbitals
(bottom row in Fig. 16) point towards each other, so that the crystal-field charges
felt by the |x2 –y2 > orbitals are more negative than those felt by the |xy> orbitals.
This yields an equal CF shift of the two bands and shifts the crossing to slightly
lower k-vectors. Second, for s || ez , which is perpendicular to the plane of the paper
in Fig. 16, Eq. (54) yields an off-diagonal spin-orbit matrix element which mixes
152 R. Skomski et al.

the bands and creates an avoided crossing near k = π/a. The gap at this degenerate
Fermi-surface crossing (DFSC), 4λ, and the derived anisotropy energy K1 (k), shown
in Fig. 17(b), are finite, in contrast to the perturbative result of Eq. (50), where the
anisotropy contribution diverges at Eo (k) = Eu (k). To appreciate this peak, is useful
to recall that typical noncubic 3d anisotropies are of the order of 0.1 meV per atom,
as compared to SOC constants λ of about 50 meV and bandwidths W in excess of
1000 meV. In other words, the avoided crossings in Fig. 17(a) may look tiny on the
scale of the bandwidth but they are huge compared to anisotropies actually realized
in solids.
The bottom panel in Fig. 17(b) shows the k-space integrated density of states as
a function of the occupancy n of the spin-down |xy> and |x2 –y2 > bands. In analogy
to Eq. (50), it is sufficient to restrict the integration to the matrix elements between
occupied (o) and unoccupied (u) states, as schematically shown in Fig. 17(a). The
anisotropy, which favors a magnetization perpendicular to the chain, also exhibits a
DFSC peak for half filling, near k = π/a, although this peak is much less pronounced
than the k-space peak.
The simple model of Fig. 16 elucidates a major aspect of itinerant anisotropy,
namely, that different pairs of 3d subbands yield positive or negative anisotropy
contributions, depending on which of the three canonical matrix elements are
realized in each magnetization direction. Including spin, this creates 10 × 10 = 100
different contributions. Each of these contributions depends on the band filling and
may further split due to the involvement of different neighbors. As a consequence,
the anisotropies exhibit a complicated oscillatory dependence on d-band filling.
Figure 18 illustrates this point for a nanoparticle with a completely filled ↑ band.

Fig. 16 Monatomic spin-chain model (top) with two orbitals per site, |xy > (center) and
|x2 –y2 > (bottom)
3 Anisotropy and Crystal Field 153

Fig. 17 Magnetic anisotropy of the spin chain of Fig. 16: (a) band structure without CF and SOC
interactions (dashed lines) and with CF and SOC interaction (solid lines) and (b) anisotropy as a
function of the electron wave vector in units of 1/a (top) and band filling of the |xy> and |x2 –
y2 > orbitals (bottom). The gray area in (a) shows the occupied states used to define the electron
count 0 ≤ n ≤ 2 in the bottom part of (b). The peaks in (b) are caused by degenerate Fermi-surface
crossing near k = π/a

Fig. 18 First-order anisotropy constant of a hexagonal nanoparticle: (a) structure and (b) tight-
binding anisotropy as a function of the number of d electrons (after Ref. 66)

By comparison, the anisotropy of rare-earth atoms in a given atomic environment


yields only two minima and two maxima, given by the quarter-shell rule of
Sect. “Rare-Earth Ions”. This simplicity originates from Hund’s rules, which yield
electron clouds of well-defined shape as a function of the number of f electrons.
In the itinerant case, each k-state corresponds to a different shape of the electron
cloud. This complicated picture starts to emerge in the simplest itinerant picture,
154 R. Skomski et al.

namely, in the diatomic pair model [60, 61], where the situation is reminiscent of a
quarter-shell rule.
It is instructive to compare the contributions of the nonperturbative DFSC
anisotropy peaks with the perturbative volume anisotropy due to Eu – Eo ∼ W. The
latter corresponds to the nearly homogeneous background in the top of Fig. 17(b)
and to the constant slopes near n = 0 and n = 2 in the bottom of Fig. 17(b).
In systems where the peak contribution is strong, a very dense k-point mesh is
necessary, or else the numerical error gets very big. The relative contribution of
the peaks depends on both the order of the anisotropy and the dimensionality of
the magnet. In one-dimensional magnets, the bulk and peak contributions to K1 are
comparable, as one may guess from the bottom of Fig. 17(b). More generally, Eq.
(53) means that perturbative anisotropy contributions scale as Km = W(λ/W)2m .
The peak contributions have a strength of λ but are restricted to a small k-
space volume of (l/W)d , so the corresponding anisotropy contribution scales as
λ(λ/W)d = W(λ/W)d + 1 . The peak contributions are therefore strongest in low-
dimensional magnets. They are of equal importance for d = 2 m – 1, that is for K1 in
one-dimensional magnets and K2 (K1 c ) in three-dimensional magnets. The latter is
fundamentally important, because it includes the anisotropy of cubic magnets such
as Fe and Ni. The former is important from a practical viewpoint, because quasi-one-
dimensional reflection from lattice planes creates pronounced peaks in the density
of states [77, 78].

First-Principle Calculations

The explanation of magnetocrystalline anisotropy by Bloch and Gentile [1] led to


the first attempt by Brooks in 1940 to describe itinerant anisotropy numerically
[53]. Early attempts to compute the anisotropy of itinerant magnets [53, 79–82]
led to substantial errors, such as wrong signs of K1 in cubic magnets. The errors are
partially due to the DFSC peaks discussed above, but they also reflect the limitations
of approximations such as tight binding. The use of self-consistent first-principle
density functional theory (DFT) has improved the situation in recent decades
[83–85], although reliable anisotropy calculations have remained a challenge,
especially for cubic magnets. Second-order anisotropy calculations for noncubic
transition-metals alloys, transition-metal contributions in rare-earth intermetallics,
and ultrathin films [86–91] are better described by DFT and have typical errors of
the order of 20–50%. However, in uniaxial magnets having nearly cubic atomic
environments, such as hcp Co, the situation is comparable to cubic magnets.
The Kohn-Sham equations, which form the basis of density functional theory,
are nonrelativistic. Spin-orbit coupling needs to be added in form of Eq. (50), which
is a second-order relativistic approximation, or a fully relativistic form, starting
from the Dirac equation. The latter is implemented in many modern codes, for
example, in the Vienna Ab Initio Simulation Package (VASP). The simplest method
to compute second-order anisotropies uses the so-called magnetic force theorem
3 Anisotropy and Crystal Field 155

[92, 93]. In this approach, the energy differences between two magnetization
directions are approximated by the difference of band-energy sums along different
magnetization directions, which can be achieved by a one-step diagonalization of
the full Hamiltonian. A better approach is to use total energy calculations, where
the energy is self-consistently calculated for each spin direction.
A specific problem is Hund’s second rule, which states that intra-atomic
electron-electron exchange favors states with large orbital momentum. The effect
is parameterized by the Racah parameter B and, in itinerant magnets, is known as
orbital polarization [89, 94]. The relative importance of this intra-atomic exchange
effect is reduced by band formation, but anisotropy calculations require a very high
accuracy, so that the corresponding orbital polarization effect cannot be ignored in
general. A simple but fairly accurate approach is to add an orbital polarization term
–½BL2 to the Hamiltonian, where B is of the order of 100 meV [94]. This term
lowers the energies of |xy> and |x2 – y2 > orbitals and enhances those of |z2 > orbitals.
The example of orbital polarization shows that correlation effects are important
in the determination of the anisotropy. In a strict sense, correlation effects involve
two or more Slater determinants [17], but sometimes their definition includes
Hund’s rule correlations. The latter are of the one-electron or independent-electron
type in the sense of a single Hartree-Fock-type Slater determinant [23]. Density
functional theory is, in principle, able to describe anisotropy, because anisotropy is
a ground-state property for any given spin direction. However, very little is known
about the density functional beyond the comfort zone of the free electron-inspired
local spin density approximation [95], including gradient corrections. For example,
rare-earth anisotropy, which is largely determined by the crystal-field interaction
of 4f charge distribution, can be cast in form of a density functional [96], but
this functional looks very different from the LSDA functional and its gradient
extensions.
One approach to approximately treat correlations is LSDA+U, where a Coulomb
repulsion parameter is added to the density functional [97]. The parameter U or, in
a somewhat more accurate interpretation, U ∗ = U –J is well-defined in the sense
that it should not be used to adjust theoretical results to achieve an agreement
with the experiment. Treating U as an adjustable parameter yields substantial
errors, of the order of 1 MJ/m3 for Ni [98]. However, similar to Hund’s-rules
correlations and LSDA, the LSDA+U approximation does not go beyond a single
Stater determinant. For example, it does not specifically address many-electron
phenomena such as spin-charge separation. The merit of the approach consists in
replacing local or quasilocal LSDA-type density functionals by density functionals
that are somewhat less inadequate for highly correlated systems. In particular, U
suppresses charge fluctuations and thereby improves the accuracy of the energy
levels connected by spin-orbit matrix elements [84]. Calculations going beyond a
single Slater determinant are still in their infancy. An analytic model calculation has
yielded Kondo-like corrections to the anisotropy [96], and dynamical mean-field
theory (DMFT) is being used to investigate the effect of charge fluctuations beyond
one-electron LSDA+U [99].
156 R. Skomski et al.

Case Studies

The magnetic anisotropies of a number of cubic and hexagonal 3d compounds are


only partially understood, both quantitatively and qualitatively. In cubic crystals, the
smallness of the anisotropy constants makes numerically calculations susceptible to
errors, for example, due to electron-electron correlations. Anisotropies in hexagonal
(and trigonal) magnets are higher, but their theoretical determination is complicated
by the fact that hexagonal crystal fields (sixfold symmetry) do not quench 3d
states (two- or fourfold symmetry). This quenching behavior is one reason for the
relatively high anisotropy of hexagonal magnets like BaFe12 O19 , SrFe12 O19 , and
YCo5 , as contrasted to tetragonal 3d magnets, such as steel. Hexagonal Co also
belongs to this high-anisotropy category, given that the atomic environment of the
Co atoms is nearly cubic.
Hexagonal ferrites. The anisotropy of Ba and Sr ferrites, which are widely
used as moderate-performance permanent magnets, is poorly understood in terms
of quantitative density-functional theory, partially due to the very narrow energy
levels. Nevertheless, early research by Fuchikami [57] traces the anisotropy to Fe
atoms on sites with a trigonal environment. An intriguing aspect of the system is
that all iron atoms in MFe12 O19 = (MO)·(Fe2 O3 )6 are ferric, Fe3+ , characterized
by half-filled 3d shells and zero anisotropy in the ground state. In more detail, the
crystal-field splitting yields an S = 5/2 ground state where two ↑ electrons occupy
a low-lying |xz> and |yz> doublet (e ), two ↑ electrons occupy an excited |xy>
and |x2 –y2 > doublet (e ), and the fifth ↑ electron occupies a |z2 > singlet (a 1 ) of
intermediate energy. The first-excited spin configuration is of the low-spin type
(S = 3/2), realized by one ↑ electron from the excited e level becoming an e
↓ electron. This spin configuration supports substantial anisotropy, because it has
odd numbers of electrons in two unquenched doublets. The splitting between the
S = 3/2 and S = 5/2 levels is fairly large (about 1 eV), but the admixiture of S = 3/2
character due to spin-orbit coupling is sufficient to create an anisotropy of the order
of 0.3 MJ/m3 .
Nickel. The anisotropies of the cubic transition metals (bcc Fe, fcc Co, fcc
Ni) have remained a moderate challenge to computational physics. Calculated
anisotropy constants are often wrong by several hundred percent and may even have
the wrong sign, that is, they predict the wrong easy axis. The choice of methods,
for example, with respect to the inapplicability of the force theorem to fourth-order
anisotropies, is one question [92, 93]. For instance, when a generalized gradient
approximation is used instead of the LSDA, the results are improved for bcc Fe
but not for Ni and Co [100]. In fact, the available choice of methods and density
functionals adds a “second-principle” component to first-principle calculations,
whose only input should be the atomic positions. Another problem is numerical
accuracy, depending on the number of k-points used.
A particularly well-investigated system is nickel [80, 82–84], where problems
are exacerbated by the smallness of the magnetic anisotropy (Table 12). The
anisotropy of Ni is determined by several contributions that largely cancel each
other: DFSC effects (Sect. “Itinerant Anisotropy”) are important, and the sum of the
3 Anisotropy and Crystal Field 157

anisotropy contributions from different orbitals and k-space regions is nearly zero. It
is also known that LSDA+U-type one-electron correlations are important in Ni. An
LSDA+U or “static DMFT” calculation was performed for Fe and Ni [84]. Values
of U* = 0.4 eV and U* = 0.7 eV have been advocated for Fe and Ni, respectively,
leading to anisotropy constants of 0.02 MJ/m3 for Fe (experiment: 0.05 MJ/m3 )
and − 0.04 MJ/m3 for Ni (experiment −0.005 MJ/m3 ). The Ni anisotropy is
overestimated, but the sign is correct, and a major reason for the correct sign is
the absence of a pocket near the X point of the fcc Brillouin zone. Without U, the
Fermi level cuts the pocket and spin-orbit matrix elements between occupied and
unoccupied states, similar to Fig. 17(a), creating an unphysical positive anisotropy
contribution.
YCo5 . The intermetallic compound YCo5 , which crystallizes in the hexagonal
CaCu5 structure, has the largest anisotropy among all know iron-series transition-
metal intermetallics, about 8 MJ/m3 at low temperature and 5 MJ/m3 at room
temperature [101]. Nearly all this anisotropy arises from the Co sublattices, in spite
of Y being a relatively heavy atom. According to Table 7, the spin-orbit coupling
of Y (32 meV) is not much smaller than that of Co (68 meV), but according to Eq.
(50), the effect of atomic SOC on the anisotropy scales is λ2 s2 , and the magnitude
of the Y spin is only about 0.3 μB , as compared to about 1.4 μB for Co [89]. In
other words, the anisotropy of YCo5 is about ten times greater than that of hcp Co,
in spite of the magnetically largely inert Y.
There are two reasons for the high anisotropy of YCo5 . First, the structure of
the YCo5 consists of alternating Co and Y-Co layers, in contrast to the nearly cubic
atomic environment in hcp Co. In this framework, the Y acts as a nonmagnetic
crystal-field source with a contribution similar to a vacuum. This has been shown in
a computer experiment [101] where the Y atoms were replaced by fictitious empty
interstices without any changes to the Co positions. The replacement reduces the
anisotropy by only 13%, which confirms that the anisotropy of YCo5 is largely due
to the anisotropic distribution of the Co atoms.
A secondary reason for the high anisotropy is that the electronic structure of
YCo5 supports a fairly strong orbital moment, about 0.2 μB per Co atom [93],
as compared to about 0.1 μB per atom in hcp Co [82]. The less quenched orbital
moment in YCo5 , which translates into enhanced anisotropy, partially reflects the
presence of degenerate |xy> and |x2 – y2 > states near the Fermi level [89]. According
to Eq. (54), the mixing of these states yields an orbital moment of up to 2 μB
per atom and a disproportionally strong anisotropy contribution (Fig. 17). More
importantly, the bands are very narrow near the Fermi level, which reduces the
denominator Eo – Eu in Eq. (50).
Iron, steel, and Fe nitride. Purified iron is magnetically very soft, but steel
formation due to the addition of carbon (Fe100–x Cx , x ≈ 4) drastically enhances
the coercivity [70, 102, 103]. The underlying physics is that carbon causes a
martensitic phase transition in bcc Fe, leading to a tetragonally distorted phase [70].
Figure 19 illustrates this mechanism, which is responsible for both the mechanical
and magnetic hardnesses of steel. The carbon occupies the octahedral interstitial
sites in the middle of the faces of the bcc unit cell (a). These octahedra are strongly
158 R. Skomski et al.

Fig. 19 Martensitic distortion of bcc Fe: (a) undistorted unit cell and (b) unit cell distorted
along the c-axis (dashed line). The martensitic distortion involves spontaneous symmetry breaking
along the a-, b-, or c-axis and extends over many interatomic distances, typically over several
micrometers


distorted: perpendicular to the faces, the Fe-Fe distance is smaller by a factor 2
than along the face diagonals. In a hard-sphere model based on an Fe radius of
1.24 Å, the radius of the interstitial site is 0.78 Å along the face diagonals but
only 0.19 Å perpendicular to the face. The atomic radius of C is about 0.77 Å
[103], so that the interstitial occupancy requires a strong tetragonal distortion. This
distortion breaks the cubic symmetry locally and, via elastic interactions between C
atoms on different interstitial sites, macroscopically. For example, 4 at% C yields an
enhancement of the c/a ratio by 3.5% [103]. Figure 19(b) shows the C occupancy
for a tetragonal distortion along the c-axis.
The martensitic lattice strain and the chemical effect due to the presence of the
carbon atoms yield almost equal uniaxial anisotropy contributions [102], and K1
is negative for Fe1-x Cx , of the order of −0.2 MJ/m3 . Cobalt addition changes the
sign of the volume magnetoelastic constant (Sect. “Magnetoelastic Anisotropy”)
and therefore the sign of the strain effect [70]. The magnetization is as high as
2.43 T in Fe65 Co35 , and the corresponding Honda steel [104] has a coercivity of
μo Hc = 0.020 T, as compared to 0.004 T for ordinary carbon steel. Such steels
dominated permanent-magnet technology in the early twentieth century and have
recently attracted renewed attention. Substantial anisotropy, K1 = 9.5 MJ/m3 , and a
magnetization of μo Ms = 1.9 T have been predicted for tetragonally distorted Fe-
Co with c/a = 1.23 [105], although such a strong distortion is virtually impossible
to sustain metallurgically. Experimental room-temperature anisotropies reach about
2.1 MJ/m3 [106] and require a large amount of Pt (about 75 vol.%).
The behavior of interstitial N in Fe is similar to that of C [107], but nitrogen has
the additional advantage of improving the magnetization in tetragonal superlattices
of Fe8 N, or Fe16 N2 [108]. It is well-established that α  -Fe16 N2 has a very high
magnetization [109, 110], about 2.8 ± 0.4 T, but the precise value has been a
subject of debate. An LSDA+U prediction of the magnetization is 2.6 T, which
3 Anisotropy and Crystal Field 159

includes a U contribution of 0.3 T [98]. Using U as an adjustable second-principle


parameter enhances the magnetization at a rate of 0.1 T/eV [111], but very large
values of U correspond to a heavy Fermion-like behavior that is contradictory to the
band structure of Fe8 N and to explicit first-principle calculations [98]. The room-
temperature K1 of the material is about 1.6 MJ/m3 [71]. LSDA and GGA reproduce
the correct order of magnitude [112].

Other Anisotropy Mechanisms

The magnetocrystalline anisotropy of Sects. Rare-Earth Anisotropy and 5 dominates


the behavior of most magnetic materials. Less commonly considered or more exotic
anisotropy mechanisms provide the leading contributions in a few systems and
substantial corrections in others. For example, two-ion anisotropies of magnetostatic
or electronic origin are usually much smaller than single-ion anisotropies, but they
dominate if the latter are zero by symmetry or by chance. An exotic mechanism
is the anisotropy of superconducting permanent magnets [113], which is not an
anisotropy in a narrow sense but reflects the interaction of local currents with the
real-structure features after field-cooling.

Magnetostatic Anisotropy

Magnetostatic dipole-dipole interaction between atomic spins yields a magneto-


static contribution to the magnetocrystalline anisotropy (MCA). Relativistically,
both spin-obit coupling and magnetostatic interactions are of the same order in the
small parameter v/c [29], but the similarities end here, and it is customary to treat
magnetostatic anisotropy contributions separately from MCA involving spin-orbit
coupling. The magnetostatic interaction energy between two dipole moments m and
m  , located at r and r  , respectively, has the form

μo 3m · R m · R − m · m R 2
EMS = (58)
4π R5

where R = r – r  . The total magnetostatic energy is obtained by summation over all


spin pairs.
In continuum theory, the summation must be replaced integration,  i ...
mi = ... M(r) dV, and it can be shown that EMS = ½μo H2 (r) dV or, equivalently

1
EMS = − μo M (r) · H (r) dV (59)
2

where H is the self-interaction field.


In a homogeneously magnetized body, the energy EMS depends on the direction
of m = m  . Figure 20 shows the “compass-needle” interpretation of this anisotropy
contribution. Neighboring spins lower their magnetostatic energy by aligning
160 R. Skomski et al.

Fig. 20 Magnetostatic contribution to the magnetocrystalline anisotropy of a layered magnet with


tetragonal symmetry. The energy of the spin configuration (a) is higher than that of (b), because
the former creates a relatively large magnetic field between the layers

parallel to the nearest-neighbor bond direction R/R, and in noncubic compounds,


this amounts to magnetic anisotropy. The corresponding contribution to K1 , which
can exceed 0.1 MJ/m3 , is especially important in some noncubic Gd-containing
magnets, because Gd combines a large atomic moment (7 μB ) with zero crystal-
field anisotropy due to its half-filled 4f shell. In cubic magnets, the anisotropy arising
from Eq. (58) is exactly zero [1], because it is a second-order anisotropy.
The anisotropy of Fig. 20 is closely related to the phenomenon of shape
anisotropy. If a homogeneously magnetized magnet has the shape of an ellipsoid,
then H(r) in Eq. (59) is also homogeneous inside the magnet (demagnetizing field).
For ellipsoids of revolution magnetized along the axis of revolution, H  = –N M,
where N is the demagnetizing factor, that is, N ≈ 0 for long needles, N = 1/3 for
spheres, and N ≈ 1 for plate-like ellipsoids [10, 114]. Turning the magnetization
 
in a direction perpendicular to the axis of revolution yields H ⊥ = – 1–2N M.
Putting H|| and H⊥ into Eq. (59) and comparing the energies EMS yields the shape
anisotropy constant:

μo  
Ksh = 1 − 3 N M2 (60)
4

This constant adds to the magnetocrystalline anisotropy constant, Keff = K1 + Ksh .


However, some precautions are necessary when using this equation. Consider a
slightly elongated magnet with N = 1/4 and zero magnetocrystalline anisotropy.
Equation (60) then predicts a positive net anisotropy constant Keff = μo M2 /16,
corresponding to a preferred magnetization direction parallel to the axis of
revolution. This is contradictory to the experiment.
In fact, the “shape anisotropy” of macroscopic magnets is merely a demag-
netizing field energy. The demagnetizing factor N is defined for uniform mag-
netization, corresponding to the Stoner-Wohlfarth model in micromagnetism, and
this nanoscale uniformity is also exploited to evaluate Eq. (59). However, in
3 Anisotropy and Crystal Field 161

Fig. 21 Micromagnetic nature of shape anisotropy in a slightly prolate but defect-free ellipsoid.
Imperfections, including nonellipsoidal edges, cause reduced nucleation fields (coercivities), which
is known as Brown’s paradox

macroscopic magnets, the magnetization state becomes nonuniform (incoherent)


due to magnetization curling [9]. The curling leads to vortex-like magnetization
states for which a shape anisotropy can no longer be meaningfully defined. Curling
reflects the strength of the magnetostatic interaction relative to the interatomic
exchange and occurs when the radius of the ellipsoid exceeds the coherence radius
Rcoh ≈ 5(A/μo Ms 2 )1/2 , or about 10 nm for a broad range of ferromagnetic materials.
Figure 21 elaborates the micromagnetic character of shape anisotropy by showing
the external nucleation field (coercivity) as a function of the particle radius.
Atomic-scale magnetism, as in Fig. 20, is realized on a sub nm length scale. On
this scale, the interatomic exchange is sufficient to ensure a parallel spin alignment,
and the magnetic anisotropy is a well-defined quantity. Elongated nanoparticles, for
example, fine-particle magnets such as Fe amalgam [13], have radii of the order
of 10 nm and are well-described by Eq. (60). Shape anisotropy is also exploited
in alnico magnets [115–118], which contain needles of high-magnetization FeCo
embedded in a nonmagnetic NiAl matrix. The radius R of the needles is smaller than
about 50 nm but substantially larger than Rcoh , which reduces the shape anisotropy
by a factor Rcoh 2 /R2 [9].

Néel’s Pair-Interaction Model

The magnetocrystalline anisotropies of Sects. “Rare-Earth Anisotropy” and 5 are


single-ion anisotropies, that is, they can be expressed in terms of atomic spin
operators such as ŝz 2 . The underlying physical phenomenon is the spin-orbit
162 R. Skomski et al.

coupling, which is separately realized in each atom and described by Eq. (21).
The single-ion mechanism does not exclude interactions between spins, such as
exchange, but the net anisotropy of a magnet is obtained by adding all single-ion
contributions. Examples of two-ion anisotropies are the magnetostatic anisotropy,
just discussed and epitomized by Eq. (58), and Néel’s phenomenological pair-
interaction model [119]. The latter uses an expansion of the anisotropy energy in
direction cosines. In the lowest order, the pair energy is equal to L (cos2 ψ – 1/3),
where L is a phenomenological parameter and ψ is the angle between bond axis and
magnetization direction. Néel’s expression is reproduced by putting m = m’ in Eq.
(58), that is, by assuming a uniform magnetization direction.
Single-ion and Néel two-ion anisotropies yield anisotropy-energy expressions of
the correct symmetry, but this does not mean that they are physically equivalent. For
example, both magnetic and nonmagnetic atoms contribute to the crystal field acting
on rare-earth ions, but the latter make no contribution in the Néel model, because is
based on pairs of magnetic atoms. Nonmagnetic ligands yield big anisotropy effects
in some materials, such as Sm2 Fe17 interstitially modified by N or C [41, 120]. The
alloy crystallizes in the rhombohedral Th2 Zn17 structure, where each Sm atom is
coordinated by three 9e interstitial sites, as shown in Fig. 22(a). The anisotropy of
Sm2 Fe17 is easy plane, that is, the Sm moment lies in the x-y-plane plane, which
also contains the 9e triangle. Heating powders of Sm2 Fe17 in N2 gas (gas-phase
interstitial modification) causes the nitrogen atoms to occupy the 9e interstices,
yielding the approximate composition Sm2 Fe17 N3 . The nitrogen addition changes
the anisotropy from easy-plane (K1 = −0.8 MJ/m3 ) to easy-axis (K1 = 8.6 MJ/m3 ),
because the virtually nonmagnetic N atoms act as strongly negative crystal-field
charges and repel the tips of the 4f charge distribution, Fig. 22(b).
One- and two-ion anisotropies are difficult to distinguish experimentally, partly
because interatomic exchange keeps neighboring spins parallel. The tempera-
ture dependence of the anisotropy is sometimes used as a criterion, scaling as
K1 (T) ∼ Ms (T)2 for magnetostatic anisotropy. However, a very similar behavior
is observed in L10 magnets such as FePt and CoPt, where the anisotropy is of the
single-ion type but requires proximity spin polarization of the 5d electrons by the
3d electrons [51, 52].

Two-Ion Anisotropies of Electronic Origin

Two-ion anisotropy is sometimes equated with magnetostatic anisotropy, but there


are also quantum-mechanical two-ion mechanisms [121]. The simplest example is
the two-ion anisotropy model described by the S = 1/2 Hamiltonian:

H = –Jxx Ŝx · Ŝx –Jyy Ŝy · Ŝy –Jzz Ŝz · Ŝz (61)

In the isotropic Heisenberg model, J xx = J yy = J zz = J , but generally J xx = J


yy J zz due to spin-orbit coupling. There is no single-ion anisotropy in the model,
=
because the operator equivalent O 2 0 (S) = 3 Sz 2 – S(S + 1) is zero for S = 1/2 and
3 Anisotropy and Crystal Field 163

Fig. 22 Anisotropy of
Sm2 Fe17 N3 : (a) interstitial
sites surrounding the Sm3+
ion in Sm2 Fe17 (blue) and (b)
change of the easy-axis
direction due to interstitial
nitrogen (yellow). Since this
anisotropy mechanism
involves one magnetic atom
only, it cannot be cast in form
of a Néel interaction

Sz = ± 1/2, but the “combined” spin S = 1, with Sz = 0 and Sz = ±1, supports


second-order anisotropy.
In the uniaxial limit, J xx = J yy = J o + J and J zz = J o – 2 J , where J o is
the isotropic Heisenberg exchange and J is relativistically small. Diagonalization
of Eq. (61) yields a singlet (S = 0) with wave function |↓↑ – ↑↓ > and energy
3J o /4, as well as triplet (S = 1). The triplet contains the Sz = ± 1 states |↑↑ > and
|↓↓>, both of energy – J o /4 – J /2, and the Sz = 0 state |↓↑ + ↑↓>, which has the
energy – J o /4 + J . Figure 23 shows the corresponding energy levels for J o > 0
and an anisotropy splitting 3 J /2 > 0. The anisotropic part of the triplet energy can
be written as

J  2 
Ea = − 3Sz − S (S + 1) (62)
2

Formally, this expression is a spin Hamiltonian in form of an operator equivalent,


but here the spin S is the combined spin of the two atoms.
The parameter J reflects spin-orbit coupling, very similar to single-
ion anisotropy and Dzyaloshinski-Moriya interactions. As emphasized in the
introduction, the Heisenberg model is isotropic, even if the bond distribution
164 R. Skomski et al.

Fig. 23 Level splitting for


the two-ion model of Eq.
(61). The anisotropic triplet is
very similar to an L = 1 or
J = 1 term in ionic
crystal-field theory, except
that the two coupled spins
reside on different ions

is anisotropic, for example, in a thin film. For example, ignoring spin-orbit


coupling and trying to explain electronic two-ion anisotropy in terms of Slater-
Pauling-Néel distance dependences yields lattice-anisotropic exchange con-
stants Jo (z – z ) = Jo (x – x ) but does not reproduce the spin-anisotropic exchange
constants Jzz (r – r ) = Jxx (r – r ) required in Eq. (61). Anisotropic exchange is
usually mixed with single-ion anisotropy and relatively small, as exemplified by
hexagonal Co, whose saturation magnetization decreases by about 0.5% on turning
the magnetization from the easy magnetization direction into the basal plane [122].
The small parameter involved is essentially K1 Vat /Jo , so that the effect can be
enhanced by reducing Jo . However, since Tc ∼ Jo , this strategy is limited to low-
temperature magnets [123, 124].
Starting from the isotropic Heisenberg model (J ), the addition of an anisotropy
term Ea ≈ –K1 Sz 2 and putting K1 = ∞ yields the classical single-ion Ising
model [125–129]. The model, which has greatly advanced the understanding of
thermodynamic phase transitions, is characterized by Sz = ±S, whereas Sx = Sy = 0
reflects the “squeezing” of quantum-mechanical degrees of freedom due to the high
anisotropy. The model requires S ≥ 1, because Eq. (52) yields zero anisotropy for
S = 1/2. However, the underlying quantum-fluctuations are ignored in classical
models in the first place, and it is common to interpret the Ising model as a classical
spin1/2 model. Quantum-mechanical Ising models are obtained by putting J xx =
J yy = 0 in Eq. (61) while allowing nonzero values of Sx and Sy , for example, in
a transverse magnetic field [130, 131]. Such two-ion models are important in the
context of quantum-phase transitions.

Dzyaloshinski-Moriya Interactions

An interaction phenomenon closely related to single-ion anisotropy, electronic pair


anisotropy, and anisotropic exchange is the Dzyaloshinski-Moriya (DM) interaction
HDM = − ½  ij Dij · Si × Sj [132–135], where i and j refer to neighboring
atoms. The DM vector Dij = − Dji reflects the local environment of the magnetic
atoms and is nonzero only in the absence of inversion symmetry. Like the spin-
3 Anisotropy and Crystal Field 165

orbit coupling, the DM interaction is derived from the Dirac equation and is of the
same order relativistically. Phenomenologically, the interaction favors noncollinear
spin states, because HDM = 0 if the spins Si and Sj are parallel. Micromagnetically,
the DM interactions can be expressed in terms of magnetization gradients ∇S and
then assume the form of Lifshitz invariants. The corresponding energy contributions
depend on the point group of the crystal or film and are zero even for some crystals
without inversion symmetry [136].
DM interactions occur in some crystalline materials, such as α-Fe2 O3
(haematite), in amorphous magnets, in spin glasses, and in magnetic nanostructures
[37, 135, 137]. The resulting canting is small, because D competes against the
dominant Heisenberg exchange J, but the canting is easily observed in hematite
and other canted antiferromagnets where there is no ferromagnetic background.
The micromagnetism of the DM interactions [138] and its competition with single-
ion anisotropy is important in the context of magnetic vortices, for example, in
MnSi [139]. The spin angles between neighboring atoms are comparable to angles
encountered in domain walls, of the order of 1◦ for material ordered at room
temperature, which reflects the common relativistic origin of both phenomena
(D in the DM interactions and K1 determining the domain-wall width). DM
noncollinearities are not be confused with noncollinearities caused by competing
Heisenberg exchange interactions.

Antiferromagnetic Anisotropy

Magnetic anisotropy is not restricted to ferromagnets, because the single-ion


mechanism is operative in each magnetic sublattice. As in ferromagnets, the
net anisotropy is obtained by adding all sublattice anisotropy contributions. The
resultant is usually nonzero; single-ion anisotropy requires a magnetic moment on
each atom, but it does not require a nonzero net magnetization. An example is CoO,
where K1 ≈ 1 MJ/m3 [78].
Antiferromagnetic anisotropy can, in principle, be extracted from the spin-flop
field. When the antiferromagnet is subjected to a sufficiently strong magnetic field
parallel to easy axis, the net magnetization jumps from zero to a finite value [129].
The corresponding spin-flop field Hsf

μo μB Hsf = 2 K1 Vat (J ∗ − K1 Vat ) (63)

reflects the competition between intersublattice exchange J * and anisotropy K1 .


Snce J *  Vat K1 in most materials, Hsf  Ha , and high fields are needed to
produce the spin-flop, even in fairly soft materials.
The anisotropy energy remains unchanged on reversing the magnetization direc-
tion, Ea (M) = Ea (−M). This means that there should be no odd-order anisotropy
contributions. However, exchange bias caused by the exchange coupling of a
ferromagnetic and an antiferromagnetic phase yields an apparent unidirectional
166 R. Skomski et al.

anisotropy on cooling through a blocking temperature that was first observed as


an asymmetric shift of the hysteresis loop by Meiklejohn and Bean 1956 [140], in
their study Co nanoparticles coated with an antiferromagnetic CoO layer. Exchange
bias may be best characterized as an inner-loop effect, caused by the external
field’s inability to overcome the high anisotropy field of the antiferromagnetic
phase.

Magnetoelastic Anisotropy

Straining a magnet with a cubic crystal structure yields a noncubic structure


with nonzero second-order magnetic anisotropy. This mechanism contributes, for
example, to the magnetic anisotropy of steel (Sect. “Case Studies”). The same
consideration applies to isotropic magnetic materials, such as amorphous and poly-
crystalline magnets, if they are rolled and extruded. However, the change in K1 is
usually very small for metallurgically sustainable strain. Magnetoelastic anisotropy
is also important in soft magnets, especially in permalloy-type materials (Fe20 Ni80 ),
where the cubic anisotropy is small and the magnetoelastic contribution, caused by
magnet processing or a substrate, often dominates the total anisotropy. Magnetoe-
lastic anisotropy is physically equivalent to magnetocrystalline anisotropy, because
a strained lattice is merely an unstrained lattice with modified atomic positions.
For example, the atomic environment in Fig. 1 can be considered as a tetragonally
strained cubic environment.
In many cases it is sufficient to describe a uniaxially strained isotropic medium
by the magnetoelastic energy:

HME λs E   E
=− 3 cos2 θ − 1 ε + ε2 − ε σ (64)
V 2 2

where σ is the uniaxial stress, ε = l/l denotes the elongation along the stress axis,
E is Young’s modulus, and θ is the angle between the magnetization and strain
axis. The strength of the magnetoelastic coupling is described by the saturation
magnetostriction λs .
Putting σ = 0 and θ = 0 and minimizing the magnetoelastic energy with respect
to ε yields the elongation ε = λs . A magnet which has a spherical shape in the
paramagnetic state becomes a prolate ferromagnet when λs > 0 but an oblate
ferromagnet when λs < 0. Physically, the spin alignment creates, via spin-orbit
coupling, an alignment of the atomic electron distributions and a change in lattice
parameters. Since λs is only 10–100 ppm in most ferromagnetic compounds, a
moderate stress σ = Eε can outweigh the spontaneous magnetostriction. This then
produces a magnetoelastic anisotropy energy density:

HME λs σ  
=− 3 cos2 θ − 1 (65)
V 2
3 Anisotropy and Crystal Field 167

and the magnetoelastic contribution to K1 KME = 3λs σ /2, which may be fairly
large.
For cubic crystallites, there are two independent magnetostriction coefficients in
the lowest order, and the polycrystalline average over all possible orientations is
[141]

2 3
λs = λ100 + λ111 (66)
5 5

where the quantities λ100 and λ111 are the spontaneous magnetostriction along
the cube edge and the cube diagonal, respectively. Experimental room-temperature
values of λs , measured in parts per million (10−6 ), are −7 for Fe, −33 for Ni,
+75 for FeCo, +40 for Fe3 O4 , −1560 for SmFe2 , and + 1800 for TbFe2 , and
practically zero for Py (permalloy, Fe20 Ni80 ) [11, 70, 115]. For example, high-
carbon steel (Fe94 C6 ) has E = 200 GPa and is strained by about 5% [103], so that
KME ≈ −0.1 MJ/m3 (see the discussion of steel magnets in Sect. “Case Studies”).

Low-Dimensional and Nanoscale Anisotropies

Nanostructuring opens a new dimension to anisotropy research and practical


applications. Surface and interface anisotropies become important on the nanoscale,
and it is possible to realize atomic environments not encountered in the bulk [9,
142]. Examples are thin films and multilayers, nanowires, single atoms, molecules,
and nanodots on surfaces, nanogranular thin-film, and bulk magnets [142]. Figure 24
shows some of these nanostructures, whose dimensions range from less than 1 nm,
for adatoms and monatomic nanowires, to 100 nm or more in nanostructured com-
posites. Most structures can be produced freestanding or deposited on substrates,
and advanced techniques are available for their fabrication and characterization (see
the other chapters of this book and Refs. [15, 143, 144]).
From a theoretical viewpoint, arbitrarily small anisotropies are important in
the theory of two-dimensional phase transitions, because they can change the
universality class from Heisenberg-like to Ising-like and even create a nonzero Curie
temperature [145, 146].

Surface Anisotropy

Surface and interface anisotropies, which are closely related, play an important role
in magnetic thin films and nanostructures. Surface anisotropies easily dominate the
bulk anisotropy in nanostructures with cubic or amorphous crystal structures, but
surface and interface contributions are also of interest in noncubic systems. For
example, L10 -ordered magnets such as FePt and CoPt can be considered as naturally
occurring multilayers. In line with other 3d anisotropies, the sign and magnitude of
surface anisotropies are difficult to predict, but some crude rules of thumb exist for
168 R. Skomski et al.

Fig. 24 Anisotropic
nanostructures: (a) thin films
(L10 -FePt/MgO), (b)
free-standing Pd zigzag
nanowire, (c) monatomic Fe
nanowire on Pt(001), and (d)
Co adatom on an insulating
substrate. First-principle
calculations often use
periodic arrays of supercells
with sufficiently big
airgaps (a)

the anisotropy as a function of band filling [60, 61]. For example, anisotropy often
changes sign between Fe and Co, the Fe preferring an easy axis perpendicular to the
Fe-Fe bonds (perpendicular to the plane).
Surface anisotropy tends to dominate when the thin-film thickness is in the range
of a few atomic layers. Phenomenologically [88, 147]

K1 = KS /t + KV (67)

where t is the film thickness, KS is the surface anisotropy, and KV includes the bulk
magnetocrystalline and shape anisotropies. Typical iron-series surface anisotropies
are of the order of 0.1–1 mJ/m2 [147], or 0.03–0.3 meV per surface transition-
metal atom, which corresponds to bulk equivalents of 0.5–5 MJ/m3 . When KV and
KS favor in-plane and perpendicular anisotropy, respectively, then there is a spin-
reorientation transition from perpendicular to in-plane as the thickness exceeds
KS /|KV |. Note that Eq. (67) does not mean that the anisotropy is limited to the
surface: the equation is asymptotic, with small contributions from subsurface atoms
and from atoms deeper in the bulk.
Thin-film, multilayer, surface, and interface anisotropies have the same physical
origin as bulk anisotropies, mostly single-ion anisotropy with magnetostatic correc-
3 Anisotropy and Crystal Field 169

Fig. 25 Effect of surface


index on the surface of bcc
Fe: (a) fourth-order in-plane
anisotropy for a (001) surface
and (b) second-order in-plane
anisotropy for a (011)
surface. Gray atoms are
subsurface atoms [148]

tions. The anisotropic distribution of exchange bonds at the surface does not create
magnetic anisotropy. The Heisenberg Hamiltonian is isotropic, even if the exchange
bonds Jij = J(ri – rj ) are anisotropic. Only relative angles between neighboring spins
matter, and the Heisenberg model is silent about the easy magnetization directions.
Both the easy axes and the strength of the anisotropy depend on the index of the
surface, and there is no reason to expect that the anisotropy axis should necessarily
be normal to the surface. For example, the (100) surface of bcc Fe, Fig. 25(a), has
fourfold in-plane symmetry and yields a fourth-order anisotropy contribution. By
comparison, the (011) surface, Fig. 25(b), has a twofold in-plane symmetry, which
yields two nonzero lowest-order anisotropy constants , K1 and K1  [148]. Surface
defects often yield substantial anisotropy contributions [88, 144]. Stepped surfaces
are an example, which can also be considered as high-index surfaces [144, 149].

Random Anisotropy in Nanoparticles, Amorphous, and Granular


Magnets

Many magnetic materials are characterized by random easy axes n(r), so that the
uniaxial anisotropy-energy expression K1 sin2 θ must be replaced by

Ha = – K1 (n · s)2 dV (68)

where s = M(r)/Ms . Atomically, K1 in Eq. (68) is the same as the K1 in Sect. “Low-
est-Order Anisotropies”, the only difference being the randomness of the local
c-axis. Random anisotropy is important in a variety of materials, including hard
and soft-magnetic polycrystalline solids [150–155], amorphous magnets [124, 137,
156], spin glasses [135], and nanoparticles [143, 157]. One example is the approach
to saturation in polycrystalline materials (Sect. “Anisotropy Measurements”).
Nanoparticles and nanoclusters are defined very similarly, but in a strict sense,
the former are random objects, whereas the latter are characterized by well-defined
atomic positions. Typical nanoparticles contain surface patches with many different
indexes, and the corresponding anisotropy contributions add.
The net anisotropy of nanoparticles is generally biaxial, involving both K1 and
K1  , and there is generally no physical justification for considering nanoparticles as
170 R. Skomski et al.

uniaxial magnets. This can be seen from Eq. (3): aside from accidental degeneracies,
there is always one axis of lowest energy. However, uniaxiality goes beyond
having an axis of lowest energy (easy axis), because it also requires the absence
of “secondary” anisotropy axes perpendicular to the easy axis. The secondary
anisotropy is important, because it causes hysteresis loops to deviate from the
uniaxial predictions.
Consider a nanoparticle with a highly disordered surface, so that each of the
NS surface atoms yields an anisotropy contribution ±Ko , where Ko is 0.03–0.3 meV
(Sect. “Surface Anisotropy”) and ± refers to orthogonal easy axes. For NS = ∞, the
surface anisotropy would average to zero, but in patches of finite NS , the averaging
is incomplete. The addition of NS random contributions √ ±Ko creates a Gaussian
distribution√of net anisotropies of the order of ±Ko / NS per surface atom [9], or
Keff = Ko NS /N averaged over all N atoms in the particle. Here the negative sign
means that the easiest axis switches into a direction perpendicular to the reference
axis (z-axis). Since Ns ∼ R2 and N ∼ R3 , Keff scales as 1/R.
Atomic-scale random-anisotropy effects in bulk solids were first discussed in
the context of amorphous magnets, which exhibit random-field [158], random-
anisotropy [159], and random-exchange spin glasses [135, 137]. In less than four
dimensions, the ground state of random-anisotropy magnets does not exhibit long-
range ferromagnetic order [135]. However, this does not preclude the use of
random-anisotropy magnets as nanostructured magnetic materials, where hysteretic
properties are important [155, 160] and true equilibrium is rarely reached. The
coercivity and remanence of atomic-scale random-anisotropy magnets were first
investigated in the late 1970s [151, 156], but a very similar situation is encountered
in nanocrystalline magnets [161, 162].
The random anisotropy in Eq. (68) creates magnetic hysteresis. In the case
of noninteracting random-anisotropy grains, which also includes noninteracting
nanoparticles,
the M(H) loops are obtained by adding the Zeeman interaction –μo
Ms H·s dV to Eq. (68), finding the M(H) loop for each direction n, and then
averaging over all n. In terms of Ha = 2 K1 /μo Ms , the behavior near remanence is
M(H) = Ms (1/2 + 2H/3Ha ). In particular, the remanence ratio Mr /Ms = M(0)/Ms
is equal to 1/2. Performing the same analysis for cubic magnets with iron-type
(K1 > 0) and nickel-type (K1 < 0) anisotropy yields the remanence ratios 0.832 and
0.866, respectively. Replacing the easy-axis anisotropy by easy-plane anisotropy
yields a very similar curve for H > 0, namely, M(H) = Ms (π/4 + H/3Ha ), and
the same asymptotic behavior (Sect. “Anisotropy Measurements”). However, the
coercive behavior is very different: random easy-axis anisotropy yields Hc = 0.479
Ha , whereas easy-plane anisotropy leads to Hc = 0.
Intergranular exchange modifies the hysteresis loops, creating some coercivity
in the easy-plane ensembles but reducing the coercivity in the case of easy-
axis anisotropy. The exchange energy density, A(∇σ m )2 , is largest for rapidly
varying magnetization directions σ m , so that exchange effects are most pronounced
grain with small radius R. In the weak-coupling limit, A/R2  K1 , there are
quantitative corrections to the hysteresis loop [9], but the strong-coupling behavior
is qualitatively different.
3 Anisotropy and Crystal Field 171

In the limit of infinite exchange, all grain magnetizations would be parallel,


σ m (r) = σ mo , and the average anisotropy of Eq. (68) would be zero by symmetry
for isotropic magnets. Large but finite exchange means that N grains are coupled
ferromagnetically, where N increases with A. Each grain yields an anisotropy
contribution ±K1 , but as in the above nanoparticle√ case, the anisotropy does not
average to zero but exhibits a distribution ±K1 / N. This yields the total energy
density:

A 1
η= 2
− K1 √ (69)
L N

where L is the magnetic correlation length, that is, the radius of the correlated
regions. In d dimensions, it is given by N = (L/R)d . Putting this expression into
Eq. (69) and minimization with respect to L yields the scaling relation

L ∼ R (δo /R)4/(4–d) (70)

where δ o = (A/K)1/2 is the domain-wall-width parameter. Equation (70) shows that


d = 4 is a marginal dimension below which small grains (R < δ o ) yield intergranular
correlations (L√> R). In three dimensions, L ∼ 1/R3 .
Since K1 / N can be considered as an effective anisotropy, the formation of
correlated regions reduces the coercivity:

Hc ∼ Ha (R/δo )2d/(4–d) (71)

In three dimensions, this means that the coercivity of random-anisotropy magnets


scales as R6 [156]. This dependence helps to reduce the coercivity of soft-magnetic
materials [155]. For example, K1 is virtually zero for amorphous alloys Fe40 Ni40 B20
and Gd25 Co75 [37]. Random anisotropy magnets having large grain sizes are in a
weak-coupling regime and exhibit high coercivities of the order of 2K1 /μo Ms , and
there is a fairly sharp transition between the strong-coupling (small R) and weak-
coupling (large R) regimes.

Giant Anisotropy in Low-Dimensional Magnets

Very high anisotropies per atom are possible in small-scale nanostructures such
as adatoms on surfaces or monatomic wires. These high anisotropies indicate
unquenched orbital moments , due to either high spin-orbit coupling or high crystal-
field symmetry. The former is realized for Co atoms on Pt(111) [163], where
a giant magnetic anisotropy of about 9 meV per Co atom has been measured.
Platinum is predisposed toward strong anisotropy, because it is close to the onset
of ferromagnetism and possesses a spin-orbit coupling of about 550 meV. A single
atom of Fe or Co easily spin-polarizes several Pt atoms, which then make large
contributions to the anisotropy. Atomically thin nanowires, such as the zigzag wire
172 R. Skomski et al.

in Fig. 24(b), may support very high anisotropy, partly due to pronounced van-Hove
peaks in the density of states. In terms of Eq. (53), van-Hove singularities near
the Fermi level correspond to small energy differences Eu – Eo . For example, an
anisotropy of 5.36 meV per atom has been predicted for free-standing ladders of Pd
atoms [164].
An upper limit to the anisotropy per atom is given by the spin-orbit coupling
constant, λ ≈ 50 meV for the late iron-series transition metals. This huge value
corresponds to 140 MJ/m3 for dense-packed atoms. It is unlikely that this anisotropy
could be exploited in nanotechnology, because anisotropy is defined as anisotropy
energy per unit volume and the requirement of isolated or freestanding wires leads to
a dilution of the anisotropy. Densification is incompatible with such huge anisotropy,
because crystal formation involves interactions of the order 1000 meV, which tend
to quench the orbital moment.
Quenching is ineffective in free-standing monatomic nanowire however, and
anisotropy energies of 20–60 meV have been predicted or experimentally inferred
for these structures. In 3d systems, anisotropies as high 6–20 meV/atom have been
calculated for free-standing linear monatomic Co wires [165]. Some monatomic
4d and 5d wires exhibit larger anisotropies, up to 60 meV per atom in stretched
Rh and Pd, respectively [166]. The high anisotropy of freestanding monatomic
nanowires indicates that some levels undergo little or no quenching. The wires
have C∞ symmetry, which leaves the states with nonzero Lz , namely, |xz> and
|yz> (Lz = ± 1) and |xy> and |x– y2 > (Lz = ± 2), completely unquenched so long
as the spin is parallel to the symmetry axis of the wire (z-axis). Figure 26 compares
the corresponding level splitting with the tetragonal one in Fig. 6. Physically, the
electrons freely orbit around the wires, because there are no in-plane crystal-field
charges that could perturb this motion. The corresponding wave functions, |xz > ± i
|yz > and |xy > ± i |x– y2 >, yield anisotropy energies of up to λ and 2λ, respectively,
depending on the number of electrons in the system.
Configurations similar to Fig. 6 also exist in a few crystalline environments.
Recent experiments have indicated that a Co ad-atom deposited on MgO shows
the giant magnetic anisotropy of 58 meV [167]. This huge anisotropy requires a
degeneracy between two levels of equal |Lz |. Co adatoms on MgO(001) have C4
symmetry. Due to Hund’s rules, the Co2+ ion (3d7 ) has one electron in the xy-xz
doublet, and this degeneracy yields a large orbital moment, <Lz > ≈ 1, and a huge
anisotropy.
The example of Co on MgO shows high anisotropy energies can also be obtained
in some crystalline environments. The C4 argument can be extended to vertically
embedded but laterally isolated wires. Such configurations might conceivably be
used for magnetic recording. In terms of thermal stability, 50 meV corresponds to
580 K, or about 2kB T per atom. For magnetic recording, one would need about 50
kB T, or chain lengths of 25 strongly exchange-coupled 3d atoms. Heavier elements
have stronger spin-orbit couplings but cannot be used for this purpose, because
3 Anisotropy and Crystal Field 173

Fig. 26 Crystal-field
splitting in an insulating
free-standing nanowire. Since
states with the same quantum
number |Lz | > 0 form
degenerate states (doublets),
there is no quenching, and
large magnetocrystalline
anisotropies K1 Vo ∼ λ are
possible

their interatomic exchange is too small to ensure ferromagnetic alignment at room


temperature. It is uncertain whether any of the approaches outlined in this section
could be used to improve areal recording densities.
Multiferroic aspects of magnetic anisotropy are an important aspect of current
research in solid-state physics and nanoscience. Electric-field control of magnetic
anisotropy in magnetic nanostructures could enable entirely new device concepts,
such as energy-efficient electric field-assisted magnetic data storage. Due to screen-
ing by conduction electrons in metals, there is no electric-field dependent bulk
anisotropy, but the surface anisotropy changes via the filling of the 3d orbitals, which
is modified by the electric field. This was demonstrated L10 -FePd and FePt thin
films immersed in a liquid electrolyte [168], where the coercivity can be modified by
an applied electric field. A common scenario is that an electric field yields a modest
change in K1 , which modifies the coercivity of the films and could be exploited for
magnetization switching [169, 170]. Similar mechanisms are realized in nanowires
on substrates, Fig. 24 [171]. For example, the application of an electric field has
been predicted to change the sign of K1 of organometallic vanadium-benzene wires
[172]. Mechanical strain and adsorbate atoms on thin films may have a similar
effect [16].

Acknowledgments This chapter has benefited from discussions with B. Balamurugan, C. Binek,
R. Choudhary, J. Cui, P. A. Dowben, A. Enders, O. Gutfleisch, G. C. Hadjipanayis, H. Herper,
X. Hong, S. S. Jaswal, P. Kharel, M. J. Kramer, P. Kumar, A. Laraoui, L. H. Lewis, S.-H. Liou,
J.-P. Liu, R. W. McCallum, O. N. Mryasov, D. Paudyal, R. Sabirianov, S. S. Sankar, T. Schrefl,
D. J. Sellmyer, J. E. Shield, A. K. Solanki, and A. Ullah. The underlying work was or has been
supported by ARO (W911NF-10-2-0099), DOE (DE-FG02-04ER46152), NSF EQUATE (OIA-
2044049), partially NSF-DMREF (1729288), HCC, and NCMN.
174 R. Skomski et al.

Appendices

Appendix A: Spherical Harmonics

Separating radial (r) and angular (θ , φ) degrees of freedom, any function f (θ , φ)


can be expanded into spherical harmonics Yl m (θ , φ). The present chapter uses this
expansion to describe (i) atomic wave functions ψ(r), as in Figs. 4 and 11, (ii) atomic
charge densities n(r), (iii) crystal-field potentials V(r) and operator equivalents O m l ,
and (iv) magnetic anisotropy energies Ea (θ , φ). These quantities differ by radial part
and physical meaning, but their angular dependences are all described by

m
Yl m (θ, φ) = Nl exp (imφ) Pl m (cos θ ) (72)

where the Pl m are the the associated Legendre polynomials. Concerning sign and
magnitude of the normalization factor N l m , we use the convention

(2l + 1) (l − m)!
Nl m = (73)
4π (l + m)!

It is sometimes useful to express Eq. (1) in terms of Cartesian coordinates or


“direction cosines” x, y, and z. Last but not least, the complex functions exp.(imφ)
may be replaced by real functions, using exp.(±imφ) = cos (mφ) ± i sin(mφ). These
real spherical harmonics, also known as tesseral harmonics, are often convenient,
because charge densities, crystal-field potentials, and anisotropy energies are real
by definition. However, the distinction remains important in quantum mechanics,
because complex and real spherical harmonics correspond to unquenched and
quenched wave functions, respectively.
A very frequently occurring function is


1 5  
Y2 0 = 3 cos2 θ –1 (74a)
2 4π

or

1 5 3z2 − r 2
Y2 0
= (74b)
2 4π r2

Note that the Cartesian coordinates require a factor 1/rl , which ensures that the
m
Yl are dimensionless and that the expansion is in terms of direction cosines x/r,
y/r, and z/r. Up to the sixth order, there are Table 16 lists real and complex spherical
harmonics up to the sixth order.
3 Anisotropy and Crystal Field 175

Table A.16 Spherical harmonics √ in several representations. For m = 0, the real representation
requires an additional factor 1/ 2, because the normalization behavior of cos(mφ) ± sin(mφ)
differs from that of exp.(imφ). Furthermore, the minus sign in N l m is not used for m = 0.
The following formulae can be used to extract the full spherical harmonics from the table:
Yl m = π-1/2 fN fP exp.(imφ), Yl m = π-1/2 fN fR /rl (m = 0), and Yl m = (2π)-1/2 fN fR /rl (m = 0).
Anisotropy energies involve even-order spherical harmonics only (gray rows)

Function fN = S ࣨm fP = Plm fR = rlYlm/ࣨ m


0
Y0 1/2 1 1
Y1 1 – 1/2 · 3/2 sinT x
Y1 0 1/2 · 3 cosT z
Y1 – 1 1/2 · 3/2 sinT y
Y2 2 1/4 · 15/2 sin2T x2 – y2
Y2 1 – 1/2 · 15/2 sinT cosT xz
Y2 0 1/4 · 5 3 cos2T – 1 3 z2 – r2
Y2 – 1 1/2 · 15/2 sinT cosT yz
Y2 – 2 1/4 · 15/2 sin2T 2 xy
Y3 3 – 1/8 · 35 sin3T x3 – 3 xy2
Y3 2 1/4 · 105/2
2
sin T cosT z (x2 – y2)
Y3 1 – 1/8 · 21 sinT  cos2T – cosT x (5 z2 – r2)
Y3 0 1/4 · 7  cos3T – 3 z (5 z2 – 3r2)
Y3 – 1 1/8 · 21 sin T  cos 2
T – cos T y (5 z2 – r2)
Y3 – 2 1/4 · 105/2
2
sin T cosT 2 xyz
Y3 – 3 1/8 · 35 sin3T 3 x2y – y3
Y4 4 3/16 · 35/2 sin4T x4 – 6 x2y2 + y4
Y4 3 – 3/8 · 35
3
sin T cosT xz (x2 – 3 y3)
Y4 2 3/8 · 5/2 sin 2
T  cos 2
T – 1 (x 2
– y2) (7 z2 – r2)
Y4 1 – 3/8 · 5
3
sinT  cos T – 3 cosT xz (7 z2 – 3 r2)
Y4 0 3/16 4
35 cos T – 30 cos T + 3 2
35 z4 – 30 z2r2 + 3 r4
Y4 – 1 3/8 · 5 sinT  cos3T – 3 cosT yz (7 z2 – 3r2)
Y4 – 2 3/8 · 5/2
2
sin T  cos T – 1 2
2 xy (7 z2 – r2)
Y4 – 3 3/8 · 35 sin3T cosT yz (3 x2 – y2)
Y4 – 4 3/16 · 35/2 sin4T 4 xy (x2 – y2)
Y5 5 – 3/32 · 77 sin5T x (x4 – 10 x2y2 + 5 y4)
Y5 4 3/16 · 385/2 sin4T cosT z (x4 – 6 x2y2 + y4)
Y5 3 – 1/32 · 385 sin 3
T  cos 2
T – 1 x (x2
– 3 y2)·(9 z2 – r2)
Y5 2 1/8 · 1155/2 sin2T  cos3T – cosT z (x2 – y2) (3 z2 – r2)
Y5 1 – 1/16 · 165/2 sinT 21 cos4T – 14 cos2T + 1) z (21 z4 – 12 z2r2 + r4)
Y5 0 1/16 · 11 63 cos5T – 70 cos3T + 15 cosT z (63 z4 – 70 z2r2 + 15 r4)
Y5 – 1 1/16 · 165/2 sinT 21 cos4T – 14 cos2T + 1) y (21 z4 – 12 z2r2 + r4)
Y5 – 2 1/8 · 1155/2 sin 2
T  cos 3
T – cos T 2 xyz (3z2 – r2)
Y5 – 3 1/32 · 385 sin3T  cos2T – 1 y (3x2 – y2) (9z2 – r2)
Y5 – 4 3/16 · 385/2 sin4T cosT 4 xyz (x2 – y2)
Y5 – 5 3/32 · 77 sin5T y (5 x4 – 10 x2y2 + y4)
Y6 6 1/64 · 3003 sin6T x6 – 15 x4y2 + 15 x2y4 – y6
Y6 5 – 3/32 · 1001 sin 5
T cos T x (x4 – 10 x2y2 + 5 y4)
Y6 4 3/32 · 91/2
4
sin T  cos T – 1 2
(x – 6 x2y2 + y4)(11 z2 – r2)
4

Y6 3 – 1/32 · 1365
3 3
sin T  cos T – 3 cosT xz (x2 – 3 y2)(11 z2 – 3 r2)
Y6 2 1/64 · 1365 sin2T 33 cos4T – 18 cos2T + 1) (x2 – y2)(33 z4 – 18 z2r2 + r4)
Y6 1 5
– 1/16 · 273/2 sinT (33 cos T – 70 cos T + 5 cosT )
3
xz (33 z4 – 30 z2r2 + 5 r4)
Y6 0 1/32 · 13 231 cos 6
T – 315 cos 4
T + 105 cos2
T – 5 231 z6 – 315 z4r2 + 105 z2r2 – 5 r6
Y6 – 1 1/16 · 273/2
5
sinT (33 cos T – 70 cos T + 5 cosT )3
yz (33z4 – 30z2r2 + 5r4)
Y6 – 2 1/64 · 1365
2 4
sin T 33 cos T – 18 cos T + 1) 2
2 xy (33 z4 – 18 z2r2 + r4)
Y6 – 3 1/32 · 1365 sin3T  cos3T – 3 cosT 3 yz (3 x2 – y2)(11 z2 – 3r2)
Y6 – 4 3/32 · 91/2 sin4T  cos2T – 1 4 xy (x2 – y2)(11 z2 – r2)
Y6 – 5 3/32 · 1001 sin5T cosT zy (5 x4 – 10 x2y2 + y4)
Y6 – 6 1/64 · 3003 sin T 6
xy (6 x4 – 20 x2y2 – 6 y4)
176 R. Skomski et al.

Appendix B: Point Groups

Table A.17 Less common space and point groups. The space groups in bold characters are
frequently encountered in magnetism and separately considered in the main text of the chapter
(Table 1)
Crystal system Point group Space group
Triclinic C1 (1) P1
Triclinic Ci (1) P1
Monoclinic C2 (2) P2, P21 , C2
Monoclinic Cs (m) Pm, Pc, Cm, Cc
Monoclinic C2h (2/m) C2/m, C2/c, P2/m, P21 /m, P2/c, P21 /c
Orthorhombic D2 (222) P222, P2221 , P21 21 2, P21 21 21 , C2221 , C222, F222, I222,
I21 21 21
Orthorhombic C2v (mm2) Pmm2, Pmc21 , Pcc2, Pma2, Pca21 , Pnc2, Pmn21 , Pba2,
Pna21 , Pnn2, Cmm2, Cmc21 , Ccc2, Amm2, Aem2, Ama2,
Aea2, Fmm2, Fdd2, Imm2, Iba2, Ima2
Orthorhombic D2h (mmm) Pnma, Pmmm, Pnnn, Pccm, Pban, Pmma, Pnna, Pmna,
Pcca, Pbam, Pccn, Pbcm, Pnnm, Pmmn, Pbcn, Pbca, Cmcm,
Cmce, Cmmm, Cccm, Cmme, Ccce, Fmmm, Fddd, Immm,
Ibam, Ibca, Imma
Tetragonal C4 (4) P4, P41 , P42 , P43 , I4, I41
Tetragonal S4 (4) P4, I4
Tetragonal C4h (4/m) P4/m, P42 /m, P4/n, P42 /n, I4/m, I41 /a
Tetragonal D4 (422) P422, P421 2, P41 22, P41 21 2, P42 22, P42 21 2, P43 22,
P43 21 2, I422, I41 22
Tetragonal C4v (4 mm) P4mm, P4bm, P42 cm, P42 nm, P4cc, P4nc, P42 mc, P42 bc,
I4mm, I4cm, I41 md, I41 cd
Tetragonal D2d (42m) P42m, P42c, P421 m, P421 c, P4m2, P4c2, P4b2, P4n2, I4m2,
I4c2, I42m, I42d
Tetragonal D4h (4/mmm) P4/mmm, P42 /mnm, I4/mmm, P4/mcc, P4/nbm, P4/nnc,
P4/mbm, P4/mnc, P4/nmm, P4/ncc, P42 /mmc, P42 /mcm,
P42 /nbc, P42 /nnm, P42 /mbc, P42 /nmc, P42 /ncm, I4/mcm,
I41 /amd, I41 /acd
Trigonal C3 (3) P3, P31 , P32 , R3
Trigonal S6 (3) P3, R3
Trigonal D3 (32) P32 12, P312, P321, P31 12, P31 21, P32 21, R32
Trigonal C3v (3 m) P3m1, P31m, P3c1, P31c, R3m, R3c
Trigonal D3d (3m) R3m, R3c, P31m, P31c, P3m1, P3c1
Hexagonal C6 (6) P6, P61 , P65 , P62 , P64 , P63
Hexagonal C3h (6) P6
Hexagonal C6h (6/m) P63 /mmc, P6/m, P63 /m
Hexagonal D6 (622) P622, P61 22, P65 22, P62 22, P64 22, P63 22
Hexagonal C6v (6 mm) P63 mc, P6mm, P6cc, P63 cm
Hexagonal D3h (6m2) P6m2, P6c2, P62m, P62c
Hexagonal D6h (6/mmm) P6/mcc, P63 /mcm
Cubic T (23) P21 3, P23, F23, I23, I21 3
(continued)
3 Anisotropy and Crystal Field 177

Table A.17 (Continued)


Crystal system Point group Space group
Cubic Td (43m) F43m, I43m, P43m, P43n, F43c, I43d
Cubic Th (m3) Pa3, Pm3, Pn3, Fm3, Fd3, Im3, Ia3
Cubic O (432) P432, P42 32, F432, F41 32, I432, P43 32, P41 32, I41 32
Cubic Oh (m3m) Fm3m, Im3m, Pm3m, Pn3m, Fd3m, Ia3d, Ia3d, Pn3n,
Pm3n, Fm3c, Fd3c

Appendix C: Hydrogen-Like Atomic 3d Wave Functions

Hydrogen-like 3d wave functions are obtained by solving the Schrödinger equation


for n = 3 (third shell) and l = 2 (d electrons). There are 2 l + 1 = 5 different orbitals,
and each can be occupied by up to two electrons. Explicitly,

|xy> = R3d (r)sin2 θ sin 2φ (75)

|x 2 − y 2 > = N R3d (r) sin2 θ cos 2φ (76)

|xz> = 2N R3d (r) sin θ cos θ cos φ (77)

 
|z2 > = R3d (r) 3 sin2 θ − 1 (78)

|yz > = 2N R3d (r) sin θ cos θ sin φ (79)


where N = 15/16π, ao = 0.529 Å, and
 
4Z 5/2 r 2 Zr
R3d (r) =  exp − (80)
81ao2 30ao3 ao

Aside from the real set of wave functions, there exist complex wave functions of
the type exp.(±imφ). The two sets of wave functions are linear combinations of
each other, and both are solutions of the Schrödinger equation. However, they are
nonequivalent with respect to orbital moment and magnetic anisotropy.
More generally, Ψ (r, φ, θ ) = Rn l (r) Yl m (φ, θ ), where it is convenient to express
the radial wave functions in terms of the parameter ro = ao /Z:
 
2 r
R1s =  exp −
ro 3 ro
178 R. Skomski et al.

   
1 r r
R2s =  2− exp −
2 2ro 3 ro 2ro

 
1 r r
R2p =  exp −
2 6ro 3 ro 2ro

   
2 r r2 r
R3s =  27 − 18 + 2 2 exp −
81 3ro 3 ro ro 3ro

   
4 r r2 r
R3p =  6 − 2 exp −
81 6ro 3 ro ro 3ro

 
4 r2 r
R3d =  exp −
81 30ro 3 ro 2 3ro

 
1 r3 r
R4f =  exp −
768 35ro 3 ro 3 4ro

From the radial wave functions, the following averages are obtained:

n2 ro 2  2 
<r 2 > = 5n + 1 − 3l (l + 1)
2

ro  2 
<r> = 3 n − l (l + 1)
2

1
<1/r> =
n2 r o

2
<1/r 2 > =
n3 ro 2 (2l + 1)

2
<1/r 3 > =
n3 ro 3 l (l + 1) (l + 2)
3 Anisotropy and Crystal Field 179

These formulae have numerous applications. For example, <r> and the square
root of <r2 > are used to estimate shell radii, <1/r> gives the electronic energy,
and <1/r3 > determines the strength of the spin-orbit coupling on which magne-
tocrystalline anisotropy relies.

References
1. Bloch, F., Gentile, G.: Zur Anisotropie der Magnetisierung ferromagnetischer Einkristalle. Z.
Phys. 70, 395–408 (1931)
2. Jäger, E., Perthel, R.: Magnetische Eigenschaften von Festkörpern. Akademie-Verlag, Berlin
(1983)
3. Buschow, K.H.J., van Diepen, A.M., de Wijn, H.W.: Crystal-field anisotropy of Sm3+ in
SmCo5 . Solid State Commun. 15, 903–906 (1974)
4. Sankar, S.G., Rao, V.U.S., Segal, E., Wallace, W.E., Frederick, W.G.D., Garrett, H.J.:
Magnetocrystalline anisotropy of SmCo5 and its interpretation on a crystal-field model. Phys.
Rev. B. 11, 435–439 (1975)
5. Cadogan, J.M., Li, H.-S., Margarian, A., Dunlop, J.B., Ryan, D.H., Collocott, S.J., Davis,
R.L.: New rare-earth intermetallic phases R3 (Fe,M)29 Xn : (R = Ce, Pr, Nd, Sm, Gd; M = Ti,
V, Cr, Mn; and X = H, N, C) (invited). J. Appl. Phys. 76, 6138–6143 (1994)
6. Wirth, S., Wolf, M., Margarian, A., Müller, K.-H.: Determination of anisotropy constants
for monoclinic ferromagnetic compounds. In: Missell, F.P., et al. (eds.) Proc. 9th Int.
Symp. Magn. Anisotropy and Coercivity in RE-TM Allyos, pp. 399–408. World Scientific,
Singapore (1996)
7. Andreev, A.V., Deryagin, A.V., Kudrevatykh, N.V., Mushnikov, N.V., Re’imer, V.A., Ter-
ent’ev, S.V.: Magnetic properties of Y2 Fe14 B and Nd2 Fe14 B and their hydrides. Zh. Eksp.
Teor. Fiz. 90, 1042–1050 (1986)
8. Kronmüller, H.: Theory of nucleation fields in inhomogeneous ferromagnets. Phys. Status
Solidi B. 144, 385–396 (1987)
9. Skomski, R.: Nanomagnetics. J. Phys. Condens. Matter. 15, R841–R896 (2003)
10. Skomski, R., Coey, J.M.D.: Permanent Magnetism. Institute of Physics, Bristol (1999)
11. Evetts, J.E. (ed.): Concise Encyclopedia of Magnetic and Superconducting Materials. Perga-
mon, Oxford (1992)
12. Sucksmith, W., Thompson, J.E.: The magnetic anisotropy of cobalt. Proc. Roy. Soc. London.
A 225, 362–375 (1954)
13. Chikazumi, S.: Physics of Magnetism. Wiley, New York (1964)
14. Thole, B.T., van der Laan, G., Sawatzky, G.A.: Strong magnetic dichroism predicted in the
M4,5 X-ray absorption spectra of magnetic rare-earth materials. Phys. Rev. Lett. 55, 2086–
2089 (1985)
15. Stöhr, J.: Exploring the microscopic origin of magnetic anisotropies with X-ray magnetic
circular dichroism (XMCD) spectroscopy. J. Magn. Magn. Mater. 200, 470–497 (1999)
16. Sander, D.: The magnetic anisotropy and spin reorientation of nanostructures and nanoscale
films. J. Phys. Condens. Matter. 16, R603–R636 (2004)
17. Ballhausen, C.J.: Ligand Field Theory. McGraw-Hill, New York (1962)
18. Bethe, H.: Termaufspaltung in Kristallen. Ann. Phys. 3, 133–208 (1929)
19. Griffith, J.S., Orgel, L.E.: Ligand-field theory. Q. Rev. Chem. Soc. 11, 381–393 (1957)
20. Newman, D.J., Ng, B.: The superposition model of crystal fields. Rep. Prog. Phys. 52, 699–
763 (1989)
21. Skomski, R.: The screened-charge model of crystal-field interaction. Philos. Mag. B. 70, 175–
189 (1994)
22. Lawrence, J.M., Riseborough, P.S., Parks, R.D.: Valence fluctuation phenomena. Rep. Prog.
Phys. 44, 1–84 (1981)
180 R. Skomski et al.

23. Fulde, P.: Electron Correlations in Molecules and Solids. Springer, Berlin (1991)
24. Reinhold, J.: Quantentheorie der Moleküle. Teubner, Stuttgart (1994)
25. Goodenough, J.B.: Magnetism and the Chemical Bond. Wiley, New York (1963)
26. Herzberg, G.: Atomic Spectra and Atomic Structure. Dover, New York (1944)
27. Pauling, L., Wilson, E.B.: Introduction to Quantum Mechanics. McGraw-Hill, New York
(1935)
28. Racah, G.: Theory of complex spectra. II. Phys. Rev. 62, 438–462 (1942)
29. Jones, W., March, N.H.: Theoretical Solid State Physics I. Wiley & Sons, London (1973)
30. Bychkov, Y.A., Rashba, E.I.: Properties of a 2D electron gas with lifted spectral degeneracy.
JETP Lett. 39, 78–81 (1984)
31. Skomski, R.: The itinerant limit of metallic anisotropy. IEEE Trans. Magn. 32(5), 4794–4796
(1996)
32. Montalti, M., Credi, A., Prodi, L., Teresa Gandolfi, M.: Handbook of Photochemistry. CRC
Press, Boca Raton (2006)
33. Taylor, K.N.R., Darby, M.I.: Physics of Rare Earth Solids. Chapman and Hall, London (1972)
34. Cole, G.M., Garrett, B.B.: Atomic and molecular spin-orbit coupling constants for 3d
transition metal ions. Inorg. Chem. 9, 1898–1902 (1970)
35. van Vleck, J.H.: On the anisotropy of cubic ferromagnetic crystal. Phys. Rev. 52, 1178–1198
(1937)
36. Skomski, R., Istomin, A.Y., Starace, A.F., Sellmyer, D.J.: Quantum entanglement of
anisotropic magnetic nanodots. Phys. Rev. A. 70, 062307 (2004)
37. Coey, J.M.: Magnetism and Magnetic Materials. University Press, Cambridge (2010)
38. Hutchings, M.T.: Point-charge calculations of energy levels of magnetic ions in crystalline
electric fields. Solid State Phys. 16, 227–273 (1964)
39. Yamada, M., Kato, H., Yamamoto, H., Nakagawa, Y.: Crystal-field analysis of the magneti-
zation process in a series of Nd2 Fe14 B-type compounds. Phys. Rev. B. 38, 620–633 (1988)
40. Herbst, J.F.: R2 Fe14 B materials: intrinsic properties and technological aspects. Rev. Mod.
Phys. 63, 819–898 (1991)
41. Coey, J.M.D., Sun, H.: Improved magnetic properties by treatment iron-based rare-earth
intermetallic compounds in ammonia. J. Magn. Magn. Mater. 87, L251–L254 (1990)
42. Duc, N.H., Hien, T.D., Givord, D., Franse, J.J.M., de Boer, F.R.: Exchange interactions in
rare-earth transition-metal compounds. J. Magn. Magn. Mater. 124, 305–311 (1993)
43. Skomski, R.: Finite-temperature behavior of anisotropic two-sublattice magnets. J. Appl.
Phys. 83, 6724–6726 (1998)
44. Akulov, N.: Zur Quantentheorie der Temperaturabhängigkeit der Magnetisierungskurve. Z.
Phys. 100, 197–202 (1936)
45. Callen, E.R.: Temperature dependence of ferromagnetic uniaxial anisotropy constants. J.
Appl. Phys. 33, 832–835 (1962)
46. Callen, H.R., Callen, E.: The present status of the temperature dependence of the magne-
tocrystalline anisotropy and the l(l+1) power law. J. Phys. Chem. Solids. 27, 1271–1285
(1966)
47. O’Handley, R.C.: Modern Magnetic Materials. Wiley, New York (2000)
48. Kronmüller, H., Fähnle, M.: Micromagnetism and the Microstructure of Ferromagnetic
Solids. University Press, Cambridge (2003)
49. Kumar, K.: RETM5 and RE2 TM17 permanent magnets development. J. Appl. Phys. 63, R13–
R57 (1988)
50. Skomski, R.: Exchange-controlled magnetic anisotropy. J. Appl. Phys. 91, 8489–8491 (2002)
51. Skomski, R., Kashyap, A., Sellmyer, D.J.: Finite-temperature anisotropy of PtCo magnets.
IEEE Trans. Magn. 39(5), 2917–2919 (2003)
52. Skomski, R., Mryasov, O.N., Zhou, J., Sellmyer, D.J.: Finite-temperature anisotropy of
magnetic alloys. J. Appl. Phys. 99, 08E916-1-4 (2006)
53. Brooks, H.: Ferromagnetic anisotropy and the itinerant electron model. Phys. Rev. 58, 909–
918 (1940)
3 Anisotropy and Crystal Field 181

54. Bruno, P.: Tight-binding approach to the orbital magnetic moment and magnetocrystalline
anisotropy of transition-metal monolayers. Phys. Rev. B. 39, 865–868 (1989)
55. Skomski, R., Kashyap, A., Enders, A.: Is the magnetic anisotropy proportional to the orbital
moment? J. Appl. Phys. 109, 07E143-1-3 (2011)
56. White, R.M.: Quantum Theory of Magnetism. McGraw-Hill, New York (1970)
57. Fuchikami, N.: Magnetic anisotropy of Magnetoplumbite BaFe12 O19 . J. Phys. Soc. Jpn. 20,
760–769 (1965)
58. Weller, D., Moser, A., Folks, L., Best, M.E., Lee, W., Toney, M.F., Schwickert, M., Thiele,
J.-U., Doerner, M.F.: High Ku materials approach to 100 Gbits/inz. TEEE Trans. Magn. 36,
10–15 (2000)
59. Brooks, M.S.S., Johansson, B.: Density functional theory of the ground-state magnetic
properties of rare earths and actinides. In: Buschow, K.H.J. (ed.) Handbook of Magnetic
Materials, vol. 7, pp. 139–230. Elsevier, Amsterdam
60. Wang, D.-S., Wu, R.-Q., Freeman, A.J.: First-principles theory of surface magnetocrystalline
anisotropy and the diatomic-pair model. Phys. Rev. B. 47, 14932–14947 (1993)
61. Skomski, R.: Magnetoelectric Néel Anisotropies. IEEE Trans. Magn. 34(4), 1207–1209
(1998)
62. Mattheiss, L.F.: Electronic structure of the 3d transition-metal monoxides: I. energy band
results. Phys. Rev. B. 5, 290–306 (1972).; “II. Interpretation”, ibidem 306–315
63. Newnham, R.E.: Properties of Materials: Anisotropy, Symmetry, Structure. University Press,
Oxford (2004)
64. Slater, J.C., Koster, G.F.: Simplified LCAO method for the periodic potential problem. Phys.
Rev. 94, 1498–1524 (1954)
65. Sutton, A.P.: Electronic Structure of Materials. University Press, Oxford (1993)
66. Skomski, R., Kashyap, A., Solanki, A., Enders, A., Sellmyer, D.J.: Magnetic anisotropy in
itinerant magnets. J. Appl. Phys. 107, 09A735-1-3 (2010)
67. Yamada, Y., Suzuki, T., Kanazawa, H., Österman, J.C.: The origin of the large perpendicular
magnetic anisotropy in Co3 Pt alloy thin films. J. Appl. Phys. 85, 5094 (1999)
68. Lewis, L.H., Pinkerton, F.E., Bordeaux, N., Mubarok, A., Poirier, E., Goldstein, J.I., Skomski,
R., Barmak, K.: De Magnete et meteorite: cosmically motivated materials. IEEE Magn. Lett.
5, 5500104-1-4 (2014)
69. Fisher, J.E., Goddard, J.: Magnetocrystalline and uniaxial anisotropy in electrodeposited
single crystal films of cobalt and nickel. J. Phys. Soc. Jpn. 25, 413–418 (1968)
70. Bozorth, R.M.: Ferromagnetism. van Nostrand, Princeton (1951)
71. Takahashi, H., Igarashi, M., Kaneko, A., Miyajima, H., Sugita, Y.: Perpendicular uniaxial
magnetic anisotropy of Fe16 N2 (001) single crystal films grown by molecular beam Epitaxy.
IEEE Trans. Magn. 35(5), 2982–2984 (1999)
72. Coene, W., Hakkens, F., Coehoorn, R., de Mooij, B.D., De Waard, C., Fidler, J., Grössinger,
R.: J. Magn. Magn. Mater. 96, 189–196 (1991)
73. Kneller, E.F., Hawig, R.: The exchange-spring magnet: a new material principle for permanent
magnets. IEEE Trans. Magn. 27, 3588–3600 (1991)
74. Klemmer, T., Hoydick, D., Okumura, H., Zhang, B., Soffa, W.A.: Magnetic hardening and
coercivity mechanisms in L10 ordered FePd Ferromagnets. Scr. Met. Mater. 33, 1793–1805
(1995)
75. Willoughby, S.D., Stern, R.A., Duplessis, R., MacLaren, J.M., McHenry, M.E., Laughlin,
D.E.: Electronic structure calculations of hexagonal and cubic phases of Co3 Pt. J. Appl. Phys.
93, 7145–7147 (2003)
76. Coey, J.M.D.: New permanent magnets: manganese compounds. J. Phys. Condens. Matter.
26, 064211-1-6 (2014)
77. Lessard, A., Moos, T.H., Hübner, W.: Magnetocrystalline anisotropy energy of transition-
metal thin films: a nonperturbative theory. Phys. Rev. B. 56, 2594–2604 (1997)
78. Skomski, R., Wei, X., Sellmyer, D.J.: Band-structure and correlation effects in the Co(111)
planes of CoO. J. Appl. Phys. 103, 07C908-1-3 (2008)
182 R. Skomski et al.

79. Fletcher, G.C.: Calculations of the first ferromagnetic anisotropy coefficient, gyromagnetic
ratio and spectroscopic splitting factor for nickel. Proc. Phys. Soc. A. 67, 505–519 (1954).
An erratum, yielding reductions of SOC constant and K1 by factors of 2 and 4, respectively,
was subsequently published as a letter to the editor: G. C. Fletcher, Proc. Phys. Soc. 78, 145
(1961)
80. Kondorski, E.M., Straube, E.: Magnetic anisotropy of nickel. Zh. Eksp. Teor. Fiz. 63, 356–365
(1972) [Sov. Phys. JETP 36, 188 (1973)]
81. Takayama, H., Bohnen, K.P., Fulde, P.: Magnetic surface anisotropy of transition metals.
Phys. Rev. B. 14, 2287–2295 (1976)
82. Trygg, J., Johansson, B., Eriksson, O., Wills, J.M.: Total energy calculation of the Mag-
netocrystalline anisotropy energy in the ferromagnetic 3d metals. Phys. Rev. Lett. 75,
2871–2875 (1995)
83. Daalderop, G.H.O., Kelly, P.J., Schuurmans, M.F.H.: First-principle calculation of the
Magnetocrystalline anisotropy energy of Iron, cobalt, and nickel. Phys. Rev. B. 41, 11919–
11937 (1990)
84. Yang, I., Savrasov, S.Y., Kotliar, G.: Importance of correlation effects on magnetic anisotropy
in Fe and Ni. Phys. Rev. Lett. 87, 216405 (2001)
85. Zhao, X., Nguyen, M.C., Zhang, W.Y., Wang, C.Z., Kramer, M.J., Sellmyer, D.J., Li, X.Z.,
Zhang, F., Ke, L.Q., Antropov, V.P., Ho, K.M.: Exploring the structural complexity of
intermetallic compounds by an adaptive genetic algorithm. Phys. Rev. Lett. 112, 045502-1-5
(2014)
86. Gay, J.G., Richter, R.: Spin anisotropy of ferromagnetic films. Phys. Rev. Lett. 56, 2728–2731
(1986)
87. Daalderop, G.H.O., Kelly, P.J., Schuurmans, M.F.H.: First-principle calculation of the
magnetic anisotropy energy of (Co)n /(X)m multilayers. Phys. Rev. B. 42, 7270–7273
(1990)
88. Johnson, M.T., Bloemen, P.J.H., den Broeder, F.J.A., de Vries, J.J.: Magnetic anisotropy in
metallic multilayers. Rep. Prog. Phys. 59, 1409–1458 (1996)
89. Daalderop, G.H.O., Kelly, P.J., Schuurmans, M.F.H.: Magnetocrystalline anisotropy of YCo5
and related RECo5 compounds. Phys. Rev. B. 53, 14415–14433 (1996)
90. Manchanda, P., Kumar, P., Kashyap, A., Lucis, M.J., Shield, J.E., Mubarok, A., Goldstein, J.,
Constantinides, S., Barmak, K., Lewis, L.-H., Sellmyer, D.J., Skomski, R.: Intrinsic properties
of Fe-substituted L10 magnets. IEEE Trans. Magn. 49, 5194–5198 (2013)
91. Kumar, P., Kashyap, A., Balamurugan, B., Shield, J.E., Sellmyer, D.J., Skomski, R.:
Permanent magnetism of intermetallic compounds between light and heavy transition-metal
elements. J. Phys. Condens. Matter. 26, 064209-1-8 (2014)
92. Wang, X., Wang, D., Wu, R., Freeman, A.J.: Validity of the force theorem for magnetocrys-
talline anisotropy. J. Magn. Magn. Mater. 159, 337–341 (1996)
93. Richter, M.: Band structure theory of magnetism in 3d-4f compounds. J. Phys. D. Appl. Phys.
31, 1017–1048 (1998)
94. Eriksson, O., Johansson, B., Albers, R.C., Boring, A.M., Brooks, M.S.S.: Orbital magnetism
in Fe, Co, and Ni. Phys. Rev. B. 42, 2707–2710 (1990)
95. Lieb, E.H.: Density functionals for coulomb systems. Int. J. Quantum Chem. 24, 243–277
(1983)
96. Skomski, R., Manchanda, P., Kashyap, A.: Correlations in rare-earth transition-metal perma-
nent magnets. J. Appl. Phys. 117, 17C740-1-4 (2015)
97. Anisimov, V.I., Zaanen, J., Andersen, O.K.: Band theory and Mott insulators: Hubbard U
instead of stoner I. Phys. Rev. B. 44, 943–954 (1991)
98. Lai, W.Y., Zheng, Q.Q., Hu, W.Y.: The giant magnetic moment and electronic correlation
effect in ferromagnetic nitride Fe16 N2 . J. Phys. Condens. Matter. 6, L259–L264 (1994)
99. Zhu, J.-X., Janoschek, M., Rosenberg, R., Ronning, F., Thompson, J.D., Torrez, M.A.,
Bauer, E.D., Batista, C.D.: LDA+DMFT approach to Magnetocrystalline anisotropy of strong
magnets. Phys. Rev. X. 4, 021027-1-7 (2014)
3 Anisotropy and Crystal Field 183

100. Freeman, A.J., Wu, R.-Q., Kima, M., Gavrilenko, V.I.: Magnetism, magnetocrystalline
anisotropy, magnetostriction and MOKE at surfaces and interfaces. J. Magn. Magn. Mater.
203, 1–5 (1999)
101. Larson, P., Mazin, I.I.: Calculation of magnetic anisotropy energy in YCo5 . J. Magn. Magn.
Mater. 264, 7–13 (2003)
102. Skomski, R., Sharma, V., Balamurugan, B., Shield, J.E., Kashyap, A., Sellmyer, D.J.:
Anisotropy of doped transition-metal magnets. In: Kobe, S., McGuinness, P. (eds.) Proc.
REPM’10, pp. 55–60. Jozef Stefan Institute, Ljubljana (2010)
103. Fast, J.D.: Gases in Metals. Macmillan, London (1976)
104. Yensen, T.D.: Development of magnetic material. Electr J. 18, 93–95 (1921)
105. Burkert, T., Nordström, L., Eriksson, O., Heinonen, O.: Giant magnetic anisotropy in
tetragonal FeCo alloys. Phys. Rev. Lett. 93, 027203-1-4 (2004)
106. Andersson, G., Burkert, T., Warnicke, P., Björck, M., Sanyal, B., Chacon, C., Zlotea, C.,
Nordström, L., Nordblad, P., Eriksson, O.: Perpendicular Magnetocrystalline anisotropy in
Tetragonally distorted Fe-Co alloys. Phys. Rev. Lett. 96, 037205-1-4 (2006)
107. Jack, K.W.: The iron—nitrogen system: the crystal structures of ε-phase iron nitrides. Act
Crystallogr. 5, 404–411 (1952)
108. The α”-Fe16 N2 unit cell contains two Fe8 N formula units. However, quoting unit cells is an
awkward notation to describe stoichiometries and rarely used. For example, one generally
writes Nd2 Fe14 B and Sm2 Co17 rather than Nd8 Fe56 B4 and Sm6 Co51
109. Coey, J.M.D., O’Donnell, K., Qinian, Q., Touchais, E., Jack, K.H.: The magnetization of
α”Fe16 N2 . J. Phys. Condens. Matter. 6, L23–L28 (1994)
110. Kim, T.K., Takahashi, M.: New magnetic material having ultrahigh magnetic moment. Appl.
Phys. Lett. 20, 492–494 (1972)
111. Ji, N., Liu, X., Wang, J.-P.: Theory of giant saturation magnetization in α”-Fe16 N2 : role of
partial localization in ferromagnetism of 3d transition metals. New J. Phys. 12, 063032-1-8
(2010)
112. Ke, L.-Q., Belashchenko, K.D., van Schilfgaarde, M., Kotani, T., Antropov, V.P.: Effects of
alloying and strain on the magnetic properties of Fe16 N2 . Phys. Rev. B. 88, 024404-1-9 (2013)
113. Trommler, S., Hänisch, J., Matias, V., Hühne, R., Reich, E., Iida, K., Haindl, S., Schultz, L.,
Holzapfel, B.: Architecture, microstructure and Jc anisotropy of highly oriented biaxially tex-
tured co-doped BaFe2 As2 on Fe/IBAD-MgO-buffered metal tapes. Supercond. Sci. Technol.
25, 84019-1-7 (2012)
114. Osborn, J.A.: Demagnetizing factors of the general ellipsoid. Phys. Rev. 67, 351–357 (1945)
115. McCurrie, R.A.: Ferromagnetic Materials—Structure and Properties. Academic Press, Lon-
don (1994)
116. Skomski, R., Liu, Y., Shield, J.E., Hadjipanayis, G.C., Sellmyer, D.J.: Permanent magnetism
of dense-packed nanostructures. J. Appl. Phys. 107, 09A739-1-3 (2010)
117. Skomski, R., Manchanda, P., Kumar, P., Balamurugan, B., Kashyap, A., Sellmyer, D.J.:
Predicting the future of permanent-magnet materials (invited). IEEE Trans. Magn. 49, 3215–
3220 (2013)
118. Zhou, L., Miller, M.K., Lu, P., Ke, L., Skomski, R., Dillon, H., Xing, Q., Palasyuk,
A., McCartney, M.R., Smith, D.J., Constantinides, S., McCallum, R.W., Anderson, I.E.,
Antropov, V., Kramer, M.J.: Architecture and magnetism of alnico. Acta Mater. 74, 224–233
(2014)
119. Néel, L.: Anisotropie Magnétique Superficielle et Surstructures d’Orientation. J. Phys.
Radium. 15, 225–239 (1954)
120. Skomski, R.: Interstitial modification. In: Coey, J.M.D. (ed.) Rare-Earth—Iron Permanent
Magnets, pp. 178–217. University Press, Oxford (1996)
121. Jensen, J., Mackintosh, A.R.: Rare Earth Magnetism: Structures and Excitations. Clarendon,
Oxford (1991)
122. Wijn, H.P.J. (ed.): Magnetic Properties of Metals: d-Elements, Alloys, and Compounds.
Springer, Berlin (1991)
184 R. Skomski et al.

123. de Jongh, L.J., Miedema, A.R.: Experiments on simple magnetic model systems. Adv. Phys.
23, 1–260 (1974)
124. Sellmyer, D.J., Nafis, S.: Phase transition behavior in a random-anisotropy system. Phys. Rev.
Lett. 57, 1173–1176 (1986)
125. Ising, E.: Beitrag zur Theorie des Ferromagnetismus. Z. Phys. 31, 253–258 (1925)
126. Brush, S.G.: History of the Lenz-Ising model. Rev. Mod. Phys. 39, 883–893 (1967)
127. Yeomans, J.M.: Statistical Mechanics of Phase Transitions. University Press, Oxford (1992)
128. Binek, C.: Ising-Type Antiferromagnets: Model Systems in Statistical Physics and in the
Magnetism of Exchange Bias Springer Tracts in Modern Physics 196. Springer, Berlin (2003)
129. Skomski, R.: Simple Models of Magnetism. University Press, Oxford (2008)
130. Sachdev, S.: Quantum Phase Transitions. University Press, Cambridge (1999)
131. Skomski, R.: On the Ising character of the quantum-phase transition in LiHoF4 . AIP Adv. 6,
055704-1-5 (2016)
132. Stevens, K.W.H.: A note on exchange interactions. Rev. Mod. Phys. 25, 166 (1953)
133. Dzyaloshinsky, I.: A thermodynamic theory of ‘weak’ ferromagnetism of antiferromagnetics.
J. Phys. Chem. Solids. 4, 241–255 (1958)
134. Moriya, T.: Anisotropic superexchange interaction and weak ferromagnetism. Phys. Rev. 120,
91–98 (1960)
135. Fischer, K.-H., Hertz, A.J.: Spin Glasses. University Press, Cambridge (1991)
136. Ullah, A., Balamurugan, B., Zhang, W., Valloppilly, S., Li, X.-Z., Pahari, R., Yue, L.-
P., Sokolov, A., Sellmyer, D.J., Skomski, R.: Crystal structure and Dzyaloshinski–Moriya
micromagnetics. IEEE Trans. Magn. 55, 7100305-1-5 (2019)
137. Moorjani, K., Coey, J.M.D.: Magnetic Glasses. Elsevier, Amsterdam (1984)
138. Skomski, R., Honolka, J., Bornemann, S., Ebert, H., Enders, A.: Dzyaloshinski–Moriya
micromagnetics of magnetic surface alloys. J. Appl. Phys. 105, 07D533-1-3 (2009)
139. Bak, P., Jensen, H.H.: Theory of helical magnetic structures and phase transitions in MnSi
and FeGe. J. Phys. C. 13, L881–L885 (1980)
140. Meiklejohn, W.H., Bean, C.P.: New magnetic anisotropy. Phys. Rev. 102, 1413–1414 (1956)
141. Kneller, E.: Ferromagnetismus. Springer, Berlin (1962)
142. Balamurugan, B., Mukherjee, P., Skomski, R., Manchanda, P., Das, B., Sellmyer, D.J.:
Magnetic nanostructuring and overcoming Brown’s paradox to realize extraordinary high-
temperature energy products. Sci. Rep. 4, 6265-1-6 (2014)
143. Balamurugan, B., Das, B., Shah, V.R., Skomski, R., Li, X.Z., Sellmyer, D.J.: Assembly of
uniaxially aligned rare-earth-free nanomagnets. Appl. Phys. Lett. 101, 122407-1-5 (2012)
144. Enders, A., Skomski, R., Honolka, J.: Magnetic surface nanostructures. J. Phys. Condens.
Matter. 22, 433001-1-32 (2010)
145. Bander, M., Mills, D.L.: Ferromagnetism of ultrathin films. Phys. Rev. B. 38, 12015–12018
(1988)
146. Shen, J., Skomski, R., Klaua, M., Jenniches, H., Manoharan, S.S., Kirschner, J.: Magnetism
in one dimension: Fe on Cu(111). Phys. Rev. B. 56, 2340–2343 (1997)
147. Gradmann, U.: Magnetism in ultrathin transition metal films. In: Buschow, K.H.J. (ed.)
Handbook of Magnetic Materials, vol. 7, pp. 1–95. Elsevier, Amsterdam (1993)
148. Sander, D., Skomski, R., Schmidthals, C., Enders, A., Kirschner, J.: Film stress and domain
wall pinning in Sesquilayer iron films on W(110). Phys. Rev. Lett. 77, 2566–2569 (1996)
149. Chuang, D.S., Ballentine, C.A., O’Handley, R.C.: Surface and step magnetic anisotropy. Phys.
Rev. B. 49, 15084–15095 (1994)
150. Coehoorn, R., de Mooij, D.B., Duchateau, J.P.W.B., Buschow, K.H.J.: Novel permanent
magnetic materials made by rapid quenching. J. Phys. 49(C-8), 669–670 (1988)
151. Callen, E., Liu, Y.J., Cullen, J.R.: Initial magnetization, remanence, and coercivity of the
random anisotropy amorphous ferromagnet. Phys. Rev. B. 16, 263–270 (1977)
152. Schneider, J., Eckert, D., Müller, K.-H., Handstein, A., Mühlbach, H., Sassik, H., Kirchmayr,
H.R.: Magnetization processes in Nd4 Fe77 B19 permanent magnetic materials. Materials Lett.
9, 201–203 (1990)
3 Anisotropy and Crystal Field 185

153. Manaf, A., Buckley, R.A., Davies, H.A.: New nanocrystalline high-remanence Nd-Fe-B
alloys by rapid solidification. J. Magn. Magn. Mater. 128, 302–306 (1993)
154. Müller, K.-H., Eckert, D., Wendhausen, P.A.P., Handstein, A., Wolf, M.: Description of
texture for permanent magnets. IEEE Trans. Magn. 30, 586–588 (1994)
155. Herzer, G.: Nanocrystalline soft magnetic materials. J. Magn. Magn. Mater. 112, 258–262
(1992)
156. Alben, R., Becker, J.J., Chi, M.C.: Random anisotropy in amorphous ferromagnets. J. Appl.
Phys. 49, 1653–1658 (1978)
157. Balasubramanian, B., Skomski, R., Li, X.-Z., Valloppilly, S.R., Shield, J.E., Hadjipanayis,
G.C., Sellmyer, D.J.: Cluster synthesis and direct ordering of rare-earth transition-metal
Nanomagnets. Nano Lett. 11, 1747–1752 (2011)
158. Imry, Y., Ma, S.-K.: Random-field instability of of the ordered state of continuos symmetry.
Phys. Rev. Lett. 35, 1399–1401 (1975)
159. Harris, R., Plischke, M., Zuckermann, M.J.: New model for amorphous magnetism. Phys.
Rev. Lett. 31, 160–162 (1973)
160. Skomski, R.: Spin-glass permanent magnets. J. Magn. Magn. Mater. 157-158, 713–714
(1996)
161. Chudnovsky, E.M., Saslow, W.M., Serota, R.A.: Ordering in ferromagnets with random
anisotropy. Phys. Rev. B. 33, 251–261 (1986)
162. Richter, J., Skomski, R.: Antiferromagnets with random anisotropy. Phys. Status Solidi B.
153, 711–719 (1989)
163. Gambaradella, P., Rusponi, S., Veronese, M., Dhesi, S.S., Grazioli, C., Dallmeyer, A., Cabria,
I., Zeller, R., Dederichs, P.H., Kern, K., Carbone, C., Brune, H.: Giant magnetic anisotropy of
single co atoms and nanoparticles. Science. 300, 1130–1133 (2003)
164. Kumar, P., Skomski, R., Manchanda, P., Kashyap, A.: Ab-initio study of anisotropy and
nonuniaxial anisotropy coefficients in Pd nanochains. Chem. Phys. Lett. 583, 109–113 (2013)
165. Dorantes-Dávila, J., Pastor, G.M.: Magnetic anisotropy of one-dimensional nanostructures of
transition metals. Phys. Rev. Lett. 81, 208–211 (1998)
166. Mokrousov, Y., Bihlmayer, G., Heinze, S., Blügel, S.: Giant Magnetocrystalline anisotropies
of 4d transition-metal Monowires. Phys. Rev. Lett. 96, 147201-1-4 (2006)
167. Rau, I.G., Baumann, S., Rusponi, S., Donati, F., Stepanow, S., Gragnaniello, L., Dreiser, J.,
Piamonteze, C., Nolting, F., Gangopadhyay, S., Albertini, O.R., Macfarlane, R.M., Lutz, C.P.,
Jones, B.A., Gamberdella, P., Heinrich, A.J., Brune, H.: Reaching the magnetic anisotropy
limit of a 3d metal atom. Science. 344, 988–992 (2014)
168. Weisheit, M., Fähler, S., Marty, A., Souche, Y., Poinsignon, C., Givord, D.: Electric field-
induced modification of magnetism in thin-film Ferromagnets. Science. 315, 349–351 (2007)
169. Shiota, Y., Maruyama, T., Nozaki, T., Shinjo, T., Shiraishi, M., Suzuki, Y.: Voltage-assisted
magnetization switching in ultrathin Fe80 Co20 alloy layers. Appl. Phys. Express. 2, 063001-
1-3 (2009)
170. Manchanda, P., Kumar, P., Fangohr, H., Sellmyer, D.J., Kashyap, A., Skomski, R.: Magne-
toelectric control of surface anisotropy and nucleation modes in L10 -CoPt thin films. IEEE
Magn. Lett. 5, 2500104-1-4 (2014)
171. Manchanda, P., Skomski, R., Prabhakar, A., Kashyap, A.: Magnetoelectric effect in Fe linear
chains on Pt (001). J. Appl. Phys. 115, 17C733-1-3 (2014)
172. Manchanda, P., Kumar, P., Skomski, R., Kashyap, A.: Magnetoelectric effect in organometal-
lic vanadium-benzene wires. Chem. Phys. Lett. 568, 121–124 (2013)
Electronic Structure: Metals and Insulators
4
Hubert Ebert, Sergiy Mankovsky , and Sebastian Wimmer

Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188
Electronic Structure Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190
Spin Density Functional Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190
Band Structure Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 196
Relativistic Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201
Adiabatic Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 204
Itinerant Magnetism of Solids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208
Stoner Model of Itinerant Magnetism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208
Slater-Pauling Curve . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 211
Heusler Alloys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
Total Electronic Energy and Magnetic Configuration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
Total Electronic Energy and Magnetic Ground State . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 218
Exchange Coupling Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220
Magneto-Crystalline Anisotropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 224
Excitations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230
Magnon Dispersion Relations Based on the Rigid Spin Approximation . . . . . . . . . . . . . . . 231
Spin Spiral Calculations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233
Excitation Spectra Based on the Dynamical Susceptibility . . . . . . . . . . . . . . . . . . . . . . . . . 237
Finite-Temperature Magnetism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 240
Methods Relying on the Rigid Spin Approximation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 240
Methods Accounting for Longitudinal Spin Fluctuations . . . . . . . . . . . . . . . . . . . . . . . . . . . 242
Coherent Treatment of Electronic Structure and Spin Statistics . . . . . . . . . . . . . . . . . . . . . . 244
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249

H. Ebert () · S. Mankovsky · S. Wimmer


München, Department Chemie, Ludwig-Maximilians-Universität, München, Germany
e-mail: Hubert.Ebert@cup.uni-muenchen.de; Sergiy.Mankovskyy@cup.uni-muenchen.de;
Sebastian.Wimmer@cup.uni-muenchen.de

© Springer Nature Switzerland AG 2021 187


J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_4
188 H. Ebert et al.

Abstract

This chapter gives an overview on the various methods used to deal with
the electronic properties of magnetic solids. This covers the treatment of
noncollinear magnetism, structural and spin disorder, as well as relativistic and
many-body effects. An introduction to the Stoner theory for itinerant or band
magnetism is followed by a number of examples with an emphasis on transition
metal-based systems. The direct connection of the total electronic energy in the
ground state and its magnetic configuration is considered next. This includes
mapping the dependence of the energy on the spin configuration on a simplified
spin Hamiltonian as provided, for example, by the Heisenberg model. Another
important issue in this context is magnetic anisotropy. As it is shown, considering
excitations from a suitable reference state provides a powerful tool to search
for stable phases, while calculating the wave vector- and frequency-dependent
susceptibility gives a sound basis to understand the dynamical properties of
magnetic solids. Finally, magnetism at finite temperature is dealt with starting
from a pure classical treatment of the problem and ending with schemes that deal
with quantum mechanics and statistics in a coherent way.

Introduction

Theory and modeling always played an important role for the understanding and
development of magnetism [1, 2, 3, 4]. An early example for this is the presence of
ring currents suggested by Ampère to explain the properties of permanent magnetic
materials. Another example is the introduction of the molecular field by Weiss when
discussing magnetism at finite temperature. Another important milestone in the the-
ory of magnetism is the Bohr-van Leeuwen theorem [3] that unambiguously made
clear that magnetism is a quantum mechanical phenomenon and for that reason
requires a corresponding description. In line with this, Heisenberg’s investigation on
the relation between the energy and the spin configuration clearly demonstrated that
spontaneous spin-magnetic ordering is connected with the exchange interaction and
is not due to the much weaker dipole-dipole interaction. Interestingly, the existence
of the electronic spin follows directly from the Dirac equation that is the proper
relativistic counterpart of Schrödinger’s equation [5]. Another important direct
consequence of Dirac’s equation is the presence of spin-orbit coupling [5] that
gives rise to many technologically important phenomena as the magnetocrystalline
anisotropy, magnetostriction [6, 7], anomalous Hall [8], and magneto-optical Kerr-
effect [9]. Although these phenomena were discovered already in the nineteenth
century, a proper explanation could be given only much later on the basis of quantum
mechanics. Spin-orbit coupling is also the origin of many other important new
effects that are exploited in the field of spintronics as, for example, the spin Hall
effect [10] and spin-orbit torque [11]. In addition, one has to mention the spin-orbit-
induced anisotropy in the magnetic exchange interaction. Apart from adding to the
magnetocrystalline anisotropy, this includes the so-called Dzyaloshinskii-Moriya
4 Electronic Structure: Metals and Insulators 189

interaction that is responsible for chiral spin configurations leading in particular to


skyrmionic spin structures [12]. Starting from a description of electron-electron
interaction in the framework of quantum electrodynamics, one is led to the Breit
interaction that can be seen as a current-current interaction [13]. This correction to
the isotropic electron-electron Coulomb interaction is anisotropic and leads to the
magnetic shape anisotropy [14].
Apart from explaining magnetic phenomena on a fundamental level in detail,
theory nowadays provides in most cases a quantitative description of these. Corre-
sponding numerical studies are in general based on a treatment of the underlying
electronic structure within the framework of spin-density functional theory [15] to
deal with electronic exchange and correlation. In spite of the many successes of
this ab initio approach to magnetism, there are several limitations. This concerns
first of all the impact of strong correlations or many-body effects that are often
discussed on the basis of simplified models or hybrid schemes as the LDA+DMFT
[16] (see section “Spin Density Functional Theory”). In addition, there are still open
questions. For example, a coherent definition of orbital magnetism [17] and with this
a prescription for its calculation were suggested only recently. Another important
field to mention in this context is magnetism at finite temperature. Although the
necessary quantum-statistical formalism [4] is available, one is in practice most
often forced to use approximate schemes. These include in particular statistical or
dynamical simulations on the basis of quasi-classical spin Hamiltonians. Ab initio
theory is still extremely helpful in this case as it provides realistic parameters. This
holds true, for example, for the exchange coupling tensor [18] and the anisotropy
constant entering the extended Heisenberg Hamiltonian or for the Gilbert damping
parameter occurring within the Landau-Lifshitz-Gilbert equation [19].
Starting from a basic knowledge in solid state theory [20], this chapter gives
an overview on the methods used to deal with the electronic properties of mag-
netic solids (section “Electronic Structure Theory”). This covers in particular
the treatment of noncollinear magnetism, structural and spin disorder, as well as
relativistic and many-body effects. An introduction to the Stoner theory for itinerant
or band magnetism is followed by a number of examples with an emphasis on
transition metal-based systems (section “Itinerant Magnetism of Solids”). The direct
connection of the total electronic energy of the ground state and its magnetic
configuration is considered next (section “Total Electronic Energy and Magnetic
Configuration”). This includes the mapping of the complex energy landscape repre-
senting the dependence on the spin configuration on a simplified spin Hamiltonian
as provided, for example, by the Heisenberg model. Another important issue of this
section will be magnetic anisotropy. As it will be shown, considering excitations
from a suitable reference state provides a very powerful tool to search for stable
phases. Calculation of the wave vector- and frequency-dependent susceptibility
provides a sound basis for understanding the dynamical properties of magnetic
solids (section “Excitations”). Finally, magnetism at finite temperature is dealt with
in the last section that presents a series of methods starting from a pure classical
treatment of the problem and leading to schemes that deal with quantum mechanics
and statistics in a coherent way (section “Finite-Temperature Magnetism”).
190 H. Ebert et al.

Electronic Structure Theory

Dealing with the electronic structure of magnetic solids requires to account for
exchange and correlation and to solve the resulting electronic structure problem.
Spin density functional theory provides a powerful and flexible platform to tackle
the first issue while still allowing for extensions aiming at an improved treatment
of correlation effects. Concerning the second issue, there are many band structure
schemes now available that provide the necessary accuracy. In particular, they allow
dealing with two important aspects that are often of crucial importance for magnetic
properties, namely, disorder and relativistic effects.

Spin Density Functional Theory

Most computational investigations on the electronic structure of magnetic materials


are nowadays based on density functional theory (DFT) [15] or extensions to it.
The major goal of this approach is to reduce the complicated many-body problem
connected with the electron system of an atom, molecule, or solid effectively to a
single-particle problem. The formal basis for this tremendous simplification is laid
by the theorems of Kohn and Hohenberg [21] that introduce the electron density
n(r) as the basic system variable:

1. The total ground state energy E of any many-body system is a functional of the
density n(r)

E[n] = F [n] + d 3 r n(r) Vext (r), (1)

where Vext (r) is an arbitrary external potential, in general the Coulomb potential
of the nuclei, and F [n] itself is a functional of the density n(r) but does not
depend on Vext (r).
2. For any many-electron system, the functional E[n] for the total energy has a
minimum equal to the ground-state energy E0 = E[n0 ] at the ground-state
density n0 (r).

Applying the variational principle to the minimal property of the energy functional
Kohn and Sham [22] derived Schrödinger-like single-particle equations whose
solution allows calculating any property of the system. For this purpose, the
functional F [n] is split into three parts:
 
n(r) n(r  )
F [n] = T [n] + 3
d r d 3r  + Exc [n], (2)
|r − r  |

with the two first terms representing the kinetic and Coulomb or Hartree energy
of the electrons. The last term is a universal functional Exc [n] that represents all
4 Electronic Structure: Metals and Insulators 191

exchange and correlation effects. Introducing an auxiliary system of noninteracting


particles with the same density n(r) as the real one, the corresponding kinetic
energy can easily be expressed leading to the formal definition of the corresponding
exchange and correlation energy functional Exc [n] as containing all remaining
many-body effects. As it is common in electronic structure theory of solids in Eq. (2)
and the following atomic Rydberg units, (h̄ = 1, me = 1/2, e2 = 2, and c = 2/α
with α the fine-structure constant) are used.
As the functional Exc [n] is universal, the resulting scheme can in principle
be applied without modification to spin-magnetic, i.e., spin-polarized, systems.
However, as the available formulations for the functional are far too complicated
to be applied to real systems, Exc [n] has to be represented in practice by a suitable
approximation. For this, it is advantageous to replace DFT by the corresponding
spin-density functional theory (SDFT) that was introduced by von Barth and Hedin
[23] and Rajagopal and Callaway [24]. Restricting to the situation of collinear
magnetism with the spin-quantization axis along the global magnetization, this leads
to the following Schrödinger-like single-particle equations:
 
−∇ 2 + Vσeff (r) φiσ (r) = iσ φiσ (r). (3)

Here φiσ (r) and iσ are the wave function and energy of the single-particle state i
with spin character σ (up or down). While these quantities have a priori no physical
meaning, they can nevertheless be used to determine the central properties of the
system. This applies in particular for the spin densities



nσ (r) = |φiσ (r)|2 , (4)
i=1

as the basic variables of the system. In Eq. (4), the summation runs over all Nσ states
with their energy iσ below the Fermi energy EF that in turn is determined by the
requirement

 
N= d 3 r n↑ (r) + n↓ (r) , (5)

where N = N↑ + N↓ is the total number of electrons. Obviously, the corresponding


particle density n(r) and spin magnetization m(r), given by the relations:

n(r) = n↑ (r) + n↓ (r) (6)


m(r) = n↑ (r) − n↓ (r), (7)

may also be chosen as alternative basic variables of a spin-polarized system.


The spin-dependent effective potential Vσeff (r) entering Eq. (3) is determined
by the requirement that the total energy E[n↑ , n↓ ] that now has to be seen as
192 H. Ebert et al.

a functional of the spin densities takes a minimum. This leads finally to the
expression:

n(r  )
Vσeff (r) = Vσext (r) + 2 d 3r  + Vσxc (r), (8)
|r − r  |

with

δExc [n↑ , n↓ ]
Vσxc (r) = . (9)
δnσ (r)

Equations (3), (8), and (9) constitute the coupled Kohn-Sham equations that
obviously have to be solved self-consistently. With this accomplished, the total
energy of the system can be obtained from:

 
Nσ  
n(r) n(r  )
E[n↑ , n↓ ] = iσ − d 3r d 3r 
|r − r  |
σ =↑,↓ i=1
 
− d 3 r Vσxc (r) nσ (r) + Exc [n↑ , n↓ ]. (10)
σ =↑,↓

When the impact of spin-orbit coupling is included into the formalism (see
section “Relativistic Effects”), the corresponding expression for E[n↑ , n↓ ] is very
convenient for investigations on the magnetic anisotropy. In this case, one can in
general neglect the change of nσ (r) with the orientation of the magnetization, and
one has to consider only the first term in Eq. (10) representing the single-particle
energies of the system.
The collinear formulation given above is adequate for most situations. In case of
noncollinear magnetic configurations with the orientation of the spin magnetization
changing with position, one has in principle to use the spin-matrix formulation
of SDFT [23]. This implies that the wave functions φiσ (r) in Eq. (3) have to be
replaced by spinors φi (r), i.e., two component wave functions, that in general will
have no pure spin character. However, very often one can still assume a uniform
orientation of the magnetization within an atomic cell. In this case the potential is
also spin-diagonal in a local frame of reference with the spin-quantization axis along
the orientation m̂ of the local spin moment m. Accordingly, one can represent the
spin-dependent part of the effective 2 × 2 potential matrix function by:

V spin (r) = σ · m̂ B(r), (11)

where σ is the vector of 2 × 2 Pauli spin matrices [5] and the effective field
B(r) = (V ↑ (r) − V ↓ (r))/2 is given by the difference of the spin-up and spin-down
potentials in the local frame. Alternatively, the spin-diagonal potential in the local
frame may be related to the 2 × 2 potential matrix function in the global frame by
4 Electronic Structure: Metals and Insulators 193

means of a transformation matrix U (θ, φ) such that: σ · m̂ = U † (θ, φ) σz U (θ, φ).


This matrix can be obtained from Eq. (80) (see below) by setting θ and φ according
to the orientation m̂ and q = 0. In either case, the form of the spin-dependent
potential again implies spinor or two-component wave functions.
As indicated, the major problem of SDFT is that the functional Exc [n↑ , n↓ ]
for the exchange-correlation is not known. A useful expression for this, that can
be justified for slowly varying densities, is supplied by the local spin-density
approximation (LSDA):

LSDA
Exc [n↑ , n↓ ] = d 3 r n(r) (n↑ (r), n↓ (r)). (12)

Here (n↑ (r), n↓ (r)) is the exchange-correlation energy per electron for the
homogeneous free electron gas with uniform spin densities n↑ (r) and n↓ (r) that can
be determined with high accuracy [25]. A fit to numerical results for (n↑ (r), n↓ (r))
allows giving explicit expressions for the corresponding spin-dependent exchange-
correlation potential Vσxc (r) = Vσxc [n↑ (r), n↓ (r)].
As an example for a spin-polarized solid, Fig. 1 gives the band structure or
dispersion relation Ej kσ and the density of states (DOS) nσ (E) of bcc-Fe as cal-
culated within LSDA (see section “Band Structure Methods”). These curves clearly
show the exchange splitting due to the spin dependency of the exchange-correlation
potential when compared with results for the corresponding paramagnetic state.
Integrating nσ (E) up to the Fermi energy EF obviously gives the number of
electrons Nσ with spin character σ . The corresponding spin-magnetic moment
M = N↑ − N↓ is given in Table 1 for bcc-Fe, fcc-Ni, and hcp-Co [27]. Obviously,

2 2

0 0
Ejkσ (eV)

-2 -2
E (eV)

-4 -4

-6 -6

-8 -8
Γ Δ X 3 2 1 0 1 2 3
wave vector k n↓(E) (sts./eV) n↑(E) (sts./eV)

Fig. 1 Dispersion relation Ej kσ (left) and density of states nσ (E) (right) of ferromagnetic bcc-Fe
as calculated within LSDA. Results for the majority (↑) and minority (↓) spin states are given in
red and blue, respectively. In addition, results for the corresponding paramagnetic state are given
in black [26]
194 H. Ebert et al.

Table 1 Spin-magnetic moment of bcc-Fe, fcc-Ni, and hcp-Co calculated on the basis of the
LSDA and the GGA, compared with experimental values. The calculated magnetic moments given
in the last two columns have been obtained for the theoretical equilibrium and experimental lattice
parameter, respectively. (All data taken from [27])
Magnetic moment (μB )
Wigner-Seitz- Bulk modulus Cohesive energy
radius (a.u.) (Mbar) (eV) (atheo ) (aexpt )
Fe LSDA 2.59 2.64 7.32 2.08 2.14
GGA 2.68 1.74 6.31 2.20 2.17
Expt. 2.67 1.68 4.28 – 2.22
Co LSDA 2.54 2.68 5.98 1.50 1.62
GGA 2.63 2.14 4.52 1.63 1.63
Expt. 2.62 1.91 4.39 – 1.72
Ni LSDA 2.53 2.50 5.45 0.59 0.61
GGA 2.63 2.08 4.18 0.65 0.63
Expt. 2.60 2.86 4.44 – 0.61

the results depend to some extent on the chosen functional for Vσxc (r) (LSDA or
GGA) and the lattice parameter. Nevertheless, the calculated moments are in fairly
good agreement with experiment.
Although LSDA turned out to be astonishingly successful for many situations,
it nevertheless shows severe limitations. For example, LSDA leads in general to an
over-binding as can be seen from the Wigner-Seitz radius given in Table 1 that is too
small when compared to experiment. Furthermore, calculations on ferromagnetic Fe
led to a lower total energy for the fcc instead of the bcc structure (see section “Total
Electronic Energy and Magnetic Ground State”). These deficits could be removed
when the generalized gradient approximation (GGA) was introduced that expresses
ExcGGA [n , n ] not only in terms of the spin densities n (r) but also of their
↑ ↓ σ
gradients ∇nσ (r). A more systematic route to derive accurate exchange-correlation
energies and corresponding potentials is supplied by the optimized potential method
that leads to a functional expressed in terms of the Kohn-Sham orbitals [15].
Unfortunately, this approach is numerically much more demanding than the very
efficient LSDA or GGA schemes in particular when a reliable representation of
correlation is incorporated. Accordingly, the development of parametrizations for
the exchange-correlation potential that are at the same time efficient and sufficiently
accurate is a field of ongoing research.
A major problem of LSDA and comparable SDFT schemes is the accurate
treatment of correlation effects in case of moderate or strong correlations as they
occur in systems with narrow energy bands. A way to cure this problem is the
use of the GW method [28] that is applicable to moderately correlated systems.
Application to Ni, for example, showed in particular a narrowing of the d-band
when compared to LSDA-based results [29] as expected from photoemission [30].
A scheme that is applicable also to strongly correlated materials at a much lower
4 Electronic Structure: Metals and Insulators 195

numerical cost is the LDA+U [31] that accounts for static correlations by adding
corrections to the LDA or LSDA Hamiltonian that depend on the Hubbard Coulomb
parameter U . In case that U is much larger than the band width W , a situation
typically met in oxides, the correction term leads in particular to a splitting into
an upper and lower Hubbard band. Dynamical correlations, on the other hand, are
accounted for by the dynamical mean field theory (DMFT) that when merged with
the LSDA leads to the combined Hamiltonian [32, 33]:

1    σσ
H = HLSDA + Umm n̂ilmσ n̂ilm σ 
2  
il mσ m σ
1   † †
− Jmm ĉilmσ ĉilm σ̄ ĉilm σ̄ ĉilmσ
2 
il mσ m

− Δl n̂ilmσ . (13)
il mσ i=id ,l=d

Here, ĉ, ĉ† , and n̂ are creation, annihilation, and particle density operators that refer
to atomic orbitals labeled by the quantum numbers l, m, and σ and site index i, with
σ σ  and J
Umm  mm corresponding to Coulomb and exchange integrals, respectively.
The quantity Δl represents the so-called double counting term that takes care that
static correlations are not accounted for twice – by the LSDA Hamiltonian HLSDA
and by its complementary DMFT counterpart HDMFT . As Eq. (13) indicates, the
DMFT correction is usually restricted to the correlated subsystem of the system as,
for example, d-states in case of transition metals (i = id , l = d). Furthermore,
the Coulomb and exchange integrals are assumed to be site diagonal (single-
site approximation) leading for the many-body problem to the same situation as
for the Anderson impurity model (AIM). Accordingly, all the various many-body
techniques available to deal with the AIM can also be used when dealing with
the combined LSDA+DMFT Hamiltonian. In most cases, Eq. (13) is dealt with
by calculating in a first step the one-electron Green function (see section “Band
Structure Methods”) associated with HLSDA . In a next step, the single-site problem
is solved by a so-called impurity solver that allows representing the impact of
HDMFT in terms of a corresponding complex and energy-dependent self-energy
ΣDMFT (E). Finally, making use of the Dyson equation (see Eq. (21) below), the
one-electron Green function of the system is updated.
Figure 2 shows typical results for the spin-dependent self-energy ΣDMFT (E) of
ferromagnetic Ni. The characteristics of these curves lead to the various correlation-
induced features expected from photoemission [30]: the real part gives rise to a
renormalization of the energy bands leading to band narrowing, reduction of the
exchange splitting, and the occurrence of a satellite structure at 6 eV binding energy.
The imaginary part, on the other hand, implies a finite lifetime of the electronic state
that increases for the d-states with distance from the Fermi energy. This is reflected
in the DOS curves by a smearing-out of its structure when compared to the LSDA
result.
196 H. Ebert et al.

Fig. 2 Left: real (red) and imaginary (blue) parts of the spin-resolved self-energy ΣDMFT (E)
of ferromagnetic Ni. Right: corresponding DOS obtained on the basis of plain LSDA and
LSDA+DMFT [26]

Band Structure Methods

Most methods for calculating the electronic structure of solids on the basis of the
Kohn-Sham equation (3) assume three-dimensional periodicity for the potential
Vσeff (r). This implies that the corresponding solutions ψj k (r) are Bloch states that
transform under a lattice translation as:

ψj k (r + R n ) = eik·R n ψj k (r) (14)

and accordingly can be labeled by the wave vector k and an additional band index
j . A direct consequence of this is that Bloch states for different wave vectors k
are orthogonal. This property simplifies the solution of the band structure problem
tremendously when using the variational principle to solve Eq. (3) and transforming
that way the problem to solving an algebraic eigenvalue problem. Constructing in
this case the basis functions such that they obey Eq. (14), one is led to a secular
equation with finite dimension for each k-vector


H k − Ej k S k α j k = 0 (15)

with the Hamilton and overlap matrices, H k and S k , referring to the basis functions
and Ej k and α j k the associated eigenvalues and eigenvectors, respectively.
Obviously one still has great freedom to construct a suitable basis set, and
accordingly there is a large number of methods and corresponding computer
codes available [34]. One route to set up an electronic structure method is to
take the tightly bound core states as frozen. This allows introducing effective
pseudo-potentials that do not show the singularity of the Coulomb potential when
4 Electronic Structure: Metals and Insulators 197

approaching the atomic nucleus. As a consequence, one may use even plane
waves as basis functions. Higher accuracy and flexibility, however, can be achieved
by using a technique called projector-augmented wave method (PAW) [35] that
mediates between pseudo-potential and so-called all-electron methods, with the
latter aiming to solve the Kohn-Sham equations directly. Again one may distinguish
between methods using analytical or numerical basis functions. Within the LMTO
method [36], for example, the Kohn-Sham equation is solved numerically at a fixed
energy Eν and angular momentum l for a spherical potential inside an atomic cell
that is approximated by a sphere. A linear combination of these solutions φlν (r)
and their energy derivatives φ̇lν (r) are augmented outside the atomic cell in a way
that leads to a decaying muffin-tin orbital that solves the Kohn-Sham equation also
within all neighboring atomic cells. The Bloch sum of such muffin-tin orbitals
is obviously a suitable basis function. By construction, it is energy independent,
but with an appropriate choice of Eν it will account, within the energy regime of
interest, for the energy dependence of the exact solution up to first order. This is a
common feature of all so-called linear methods [36] like the LAPW [36, 37]. The
special choice of the basis function of the LMTO method has the additional feature
that it is minimal: this means that one can restrict the angular momentum expansion
of the basis function in line with chemical intuition, i.e., for transition metals, one
should go at least up to d-states with l = 2. Solving the resulting secular equation
or algebraic eigenvalue problem, one gets finally the energy eigenvalue Ej k of the
Bloch states together with a corresponding representation of their wave functions
ψj k (r) in terms of the radial functions φlν (r) and their energy derivatives φ̇lν (r):
 jk jk
ψj k (r) = Alm φlν (r) Ylm (r̂) + Blm φ̇lν (r) Ylm (r̂), (16)
lm
jk jk
where the expansion coefficients Alm and Blm are given by the eigenvectors α j k
and Ylm (r̂) are spherical harmonics. A similar representation of the Bloch wave
functions is obtained for most other all-electron band structure methods.
Although most band structure methods are formulated as k-space methods
assuming three-dimensional periodicity, they can nevertheless be applied to situ-
ations with lower dimensionality or symmetry by means of the super-cell technique.
This is illustrated for the case of a random substitutionally disordered binary alloy
Ax B1−x in Fig. 3. Instead of dealing with an, in principle, infinitely large unit cell
that represents the random distribution of the A- and B-atoms on the geometric
lattice, periodic boundary conditions are imposed implying the use of a finite size
super-cell that is enlarged when compared to the unit cell of the underlying lattice.
To achieve a reliable representation of the configurational average of a disordered
alloy, obviously the unit cell has to be large enough and a representative average has
to be taken concerning the atomic configuration within the super-cell [38] using,
for example, the concept of a special quasi-random structure [39]. As indicated by
the lower panel of Fig. 3, the same type of reasoning can be applied when dealing
with the problem of a disordered spin configuration of a solid that may occur due to
thermal spin fluctuations [40]. The super-cell technique is also frequently applied
198 H. Ebert et al.

Fig. 3 Top row: representation of a random substitutionally disordered binary alloy Ax B1−x (left)
by means of the super-cell technique (middle) and by means of an effective medium theory (right).
Bottom row: corresponding application of these schemes to the problem of a disordered spin
configuration

in case of reduced dimensionality of the system as, for example, when dealing with
surfaces or impurities in a host system. In the latter case obviously the size of the
super-cell has to be chosen large enough to avoid an interaction of impurities in
neighboring super-cells.
Instead of representing the electronic structure of a solid in terms of Bloch
states with associated wave functions ψj k (r) and energy eigenvalues Ej k , this can
be done by means of the corresponding retarded single-particle Green function
G+ (r, r  , E). With the solutions to the band structure problem available, the
Green function G+ (r, r  , E) can be given via the so-called Lehmann spectral
representation [41]

 ψj k (r) ψj†k (r  )
+ 
G (r, r , E) = lim , (17)
→0 E − Ej k + i
jk

that allows straightforwardly to derive convenient expressions, for example, for the
density of states n(E) and electron density n(r), respectively:

1
n(E) = −  d 3 r G+ (r, r, E) (18)
π
4 Electronic Structure: Metals and Insulators 199

 EF
1
n(r) = −  dE G+ (r, r, E). (19)
π

Although Eq. (17) is frequently used, it is not very convenient as one needs in
principle the whole spectrum connected with the underlying electronic Hamiltonian
to get the Green function for a given energy E. An alternative to this is offered
by the multiple scattering theory-based KKR (Korringa-Kohn-Rostoker) formalism
that leads to the following expression for G+ (r, r  , E) [42]:
  
i× 
G+ (r, r  , E) = ZLi (r, E) τLL
ii
 (E) ZL (r , E)
L,L

−δii  ZLi (r < , E) JLi× (r > , E) (20)
L

that in particular does not require Bloch translational symmetry for the system
considered. The functions ZLi (r, E) and JLi (r, E) in Eq. (20) are regular and
irregular solutions, respectively, to the Kohn-Sham equation with angular character
L = (l, m) for r in the atomic cell at site i and a specific normalization [42]. The
quantity τLLii  (E) is the so-called scattering path operator that transfers a wave with

character L coming in at site i  into a wave going out from site i with character L
with all intermediate scattering events accounted for. From Eq. (18), it is obvious
that the site-diagonal quantity τLL ii (E) determines in particular the variation of the

density of states ni (E) at site i with energy E.
The use of the Green function offers many advantages when dealing with
embedded subsystems, response functions, spectroscopy, disorder, or the many-
body problem. To a large extent, this is due to the Dyson equation that allows to
express the Green function G+ (r, r  , E) of a complex system on the basis of that of
a simpler reference system (G+ 
0 (r, r , E)) and the arbitrary perturbing Hamiltonian
Hpert (r) that connects the two systems:

G+ (r, r  , E) = G+ 
0 (r, r , E) + d3 r  G+ 
0 (r, r , E)
Ω

Hpert (r  ) G+ (r  , r  , E), (21)

with Ω the region for which Hpert (r) has to be accounted for. For a substitutional
impurity, this would include the atomic cell of the impurity and the region of the
neighboring host atoms that are distorted by the impurity.
The Green function formalism is particularly useful when dealing with the
electronic structure of disordered systems. By using the concept of the molecular
field, Soven [43] introduced the Coherent Potential Approximation (CPA) approach
when dealing with disordered substitutional alloys. The corresponding hypothetical
effective CPA medium plays the role of the molecular field and is constructed such
that it represents the configurational average for the alloy as accurate as possible
200 H. Ebert et al.

(see Fig. 3). The standard CPA is a so-called single-site theory implying that the
occupation of neighboring lattice sites is uncorrelated, i.e., short-range order is
excluded. Within the KKR approach, the CPA medium is therefore determined by
requiring that for an Ax B1−x alloy, the embedding of an A- or B-atom into the CPA
medium should on the average lead to no additional scattering:

A + (1 − x) τ B = τ CPA .
x τ ii ii ii
(22)

Here the component-projected scattering path operators τ ii


α represent the single-site
embedding of the component α into the CPA medium according to Eq. (21). When
using these quantities together with the corresponding component-related wave
functions in Eq. (20), one gets obviously access to component-specific properties as
the partial DOS of an alloy. Corresponding results for the disordered ferromagnetic
alloy fcc-Ni0.8 Pd0.2 are shown in Fig. 4. The left column gives the spin-resolved
band structure in terms of the Bloch spectral function AB σ (k, E) that can be seen
as the Fourier transform of the real space Green function G+ 
σ (r, r , E), while the

2 2 2
0 0 0
-2 -2 -2
E (eV)

E (eV)

E (eV)
-4 -4 -4
-6 -6 -6
-8 -8 Ni Pd -8
Γ Δ X 0 0.5 1 1.5 2 2.5 0 0.5 1 1.5 2
wave vector k n↑(E) (sts./eV) n↑(E) (sts./eV)
2 2 2
0 0 0
-2 -2 -2
E (eV)

E (eV)

E (eV)

-4 -4 -4
-6 -6 -6
-8 -8 Ni Pd -8
Γ Δ X 0 0.5 1 1.5 2 2.5 0 0.5 1 1.5 2
wave vector k n↓(E) (sts./eV) n↓(E) (sts./eV)

Fig. 4 Left: spin-resolved Bloch spectral function AB σ (k, E) of the disordered ferromagnetic alloy
fcc-Ni0.8 Pd0.2 calculated on the basis of the CPA. Middle and right column: corresponding spin-
resolved partial density of states nασ (E) for α = Ni or Pd, respectively. The top and bottom row
give results for spin up and down, respectively. As a reference, the dispersion relation Ej kσ of pure
Ni is superimposed as a black line to AB Ni
σ (k, E) (left). In addition, nσ (E) for pure ferromagnetic
Pd
Ni (middle) and of nσ (E) for pure paramagnetic Pd (right) are included in the figures as dashed
lines [26]
4 Electronic Structure: Metals and Insulators 201

middle and right columns give the spin-resolved partial density of states nασ (E) for
α = Ni or Pd, respectively. Comparison of AB σ (k, E) with the dispersion relation
Ej kσ of fcc-Ni and of nNiσ (E) for the alloy with that for pure Ni clearly shows the
smearing-out of these curves for the alloy in particular in the regime of the d-states.
This reflects the fact that for the alloy the wave vector k is not a good quantum
number. In fact the width of the AB σ (k, E) functions can be interpreted as a measure
for the electronic lifetime to be used in the calculation of the residual resistivity on
the basis of the Boltzmann formalism [44]. The right column of Fig. 4 gives results
for the partial DOS nPd σ (E) of Pd in fcc-Ni0.8 Pd0.2 and for pure paramagnetic Pd
that clearly shows that Pd gets spin-polarized due to hybridization leading to an
induced spin-magnetic moment of about 0.21 μB .
It is important to note that the concept of the CPA is not restricted to alloys but
can be applied to any type of disorder. Making use of the alloy analogy, the CPA is
used, for example, within the disordered local moment (DLM) model [45] to per-
form an average over spin configurations connected with thermal spin fluctuations
(see section “Coherent Treatment of Electronic Structure and Spin Statistics”).

Relativistic Effects

A most coherent way to account for the influence of relativistic effects is to work on
the basis of the Dirac Hamiltonian [5]:

1 2
ĤD = −icα · ∇ + c (β − 1) + V̄ (r) + β σ · B(r) + eα · A(r). (23)
2

Here c is the speed of light, αi and β are the standard 4 × 4 Dirac matrices, cα
is the electronic velocity operator, σi are the 4 × 4 spin matrices, and the local
potential may involve a spin-independent part V̄ (r), an effective magnetic field
(B(r)) coupling only to the spin [46] and the vector potential (A(r)) coupling
to the electronic current, where the effective fields B(r) and A(r) combine
exchange-correlation and possible external contributions. This approach implies a
four-component electronic wave function ψ(r, E) or bi-spinor, respectively, with a
large and small component [5]. Corresponding versions based on this framework
have been worked out for several band structure methods [47, 48] assuming in
general a spin-dependent potential only (A(r) = 0). The appealing feature of these
schemes is that they treat all relativistic effects and magnetic ordering on the same
footing.
Alternatively, one may apply a Foldy-Wouthuysen transformation [5] to the Dirac
equation. Considering only a spherical scalar potential V (r) in Eq. (23) and keeping
only terms up to the order of 1/c2 , this leads to three relativistic corrections when
compared to the Schrödinger Hamiltonian [49]:

1 4
Hmass = −β p (24)
c2
202 H. Ebert et al.

1 2
HDarwin = ∇ V (r) (25)
2c2
1 1 ∂V (r)
HSOC = 2 σ · l. (26)
c r ∂r

The first two corrections, mass enhancement (Hmass ) and Darwin (HDarwin ) terms,
do not involve the spin operator σ , and for that reason, they are called scalar
relativistic. In general, these corrections give rise to a downward shift in energy
for s- and p-states and, in response to this, to an upward shift in energy for d-states
[50]. For a paramagnetic metal, this influences the density of states at the Fermi
level and this way the tendency toward spontaneous formation of spin magnetism
via the Stoner mechanism (see section “Stoner Model of Itinerant Magnetism”).
The third correction term is the spin-orbit coupling (HSOC ), which is given in its
commonly simplified form, that holds in case of a spherically symmetric scalar
potential V (r). As HSOC couples the electronic orbital and spin degrees of freedom,
this term will lead to the removal of degeneracies for any material. For a spin-
polarized material, it leads furthermore to a reduction in symmetry as reflected by
many important properties as, for example, the magnetocrystalline anisotropy [51,6]
or galvanomagnetic [8] and magneto-optical [9] properties.
In practice, most electronic structure calculations are based on relativistic
correction schemes that first of all aim to eliminate the small component from
the Dirac equation leading to a two-component formalism [52, 53]. Within so-
called scalar relativistic calculations, spin-orbit coupling is ignored leading to no
technical changes when compared to a nonrelativistic calculation on the basis of
the LSDA that treat spin-up and spin-down states separately. However, including
HSOC does not allow this simplification any more requiring higher computational
effort. While HSOC can still be accounted for when setting up the basis functions of
a band structure scheme [52, 54], it is in most cases seen as a correction to a scalar
relativistic Hamiltonian and therefore accounted for only in the variational step of a
standard band structure scheme [36, 37].
As an illustration, Fig. 5 gives the dispersion relation Ej k of ferromagnetic Ni
calculated in a non-, scalar, and fully relativistic way. The left panel of the figure
shows the impact of the scalar relativistic corrections Hmass and HDarwin . The
middle panel shows results that account in addition for the spin-diagonal part of
HSOC proportional to lz σz [55]. Although this term leaves spin as a good quantum
number, it obviously removes many degeneracies and band crossings giving rise, for
example, to spin-orbit-induced orbital magnetic moments [55]. Accounting finally
for the full spin-orbit coupling HSOC (right panel), one finds further removals of
band crossings and a coupling of the two spin systems. This spin mixing gives rise
to so-called hot spots in the band structure [56] that are important, for example, for
spin-flip relaxation processes.
The relativistic corrections mentioned so far concern only the kinetic part of
the electronic Hamiltonian, but not the effective potential. Starting from a fully
relativistic framework in fact corresponding corrections have to be expected when
4 Electronic Structure: Metals and Insulators 203

-0.5 -0.5

-1 -1
Ejkσ (eV)

-1.5 -1.5

-2 -2
NREL SOC zz FREL

-2.5 -2.5
Γ Δ X Γ Δ X Γ Δ X
wave vector k

Fig. 5 Dispersion relation Ej k of ferromagnetic Ni. Left: spin-resolved curves Ej kσ for non-
(black lines) and scalar (up and down triangles) relativistic mode. Middle: relativistic mode that
accounts only of the spin-diagonal part of HSOC proportional to lz σz (black lines) and scalar
relativistic mode (up and down triangles). Right: fully (black lines) and scalar (up and down
triangles) relativistic mode [26]

considering the free electron gas as a reference system [15]. The impact of such
corrections to the exchange-correlation potential has been monitored in particular
for paramagnetic materials [57]. For magnetic materials, a coherent derivation
of nonrelativistic spin-density functional theory could be given by starting from
a relativistic formalism [24]. Later on, a relativistic formulation of spin-density
functional theory assuming a spin-dependent exchange-correlation potential that
couples only to the spin degree of freedom was worked out [58, 46] leading
to Eq. (23) with A(r) = 0. This simplification seems to be acceptable for
many magnetically ordered transition metal systems. On the other hand, spin-orbit
coupling gives for spin-polarized materials automatically rise to orbital magnetism
[59] that can be associated with a corresponding orbital current. Accordingly,
relativistic spin-density functional theory should in principle be replaced by a
current density formalism with the four current as a basic system variable [15],
i.e., one has B(r) = 0 and A(r) = 0 in Eq. (23). While various developments have
been made in this direction [60], there are no functionals for the corresponding
exchange-correlation available at the moment. As a consequence, the feedback of
orbital magnetism or polarization on the electronic Hamiltonian has been accounted
for primarily by means of hybrid schemes like the OP- [61], the LSDA+U- [62], or
the LSDA+DMFT-formalism [63].
Another consequence of a fully relativistic formalism is a modification of the
Coulomb potential and the occurrence of the Breit interaction [13]. The latter one
can be seen as a current-current interaction and accordingly can be represented
by a corresponding vector potential A(r) in Eq. (23). For magnetic materials, the
Breit interaction contributes not only to the total energy but also to its magnetic
anisotropy. In fact it has been pointed out that the Breit interaction is the quantum
204 H. Ebert et al.

mechanical source for the classical dipole-dipole interaction giving rise to the
magnetic shape anisotropy [14].

Adiabatic Dynamics

The features of the electronic band structure determine not only the ground-state
properties of a material but also the dynamical properties of electrons in the presence
of external time-dependent perturbations. When a perturbation varies slowly in time,
the dynamics of the electronic subsystem can be efficiently described using the
concept of the Berry phase [64] arising during the adiabatic evolution of electronic
quantum states. According to its definition, the Berry phase is connected to a
parameter-dependent Hamiltonian [65,66,67]. For systems with a periodic effective
potential Veff (r + R n ) = Veff (r) with R n a lattice translation vector, leading to a
Bloch-like solution of the eigenvalue problem (see Eq. (14)):

ψj k (r) = eik·r uj k (r), (27)

2

the unitary transformation of the Hamiltonian H = 2m + Veff (r) leads to a
momentum-dependent Hamiltonian (for details see [67])

(p̂ + h̄k)2
H(k) = e−ik·r Heik·r = + Veff (r). (28)
2m

This allows treating the Brillouin zone as the parameter space of the transformed
Hamiltonian H(k) with cell-periodic eigenfunctions uj k (r). The Berry phase is
expressed as an integral over the path C in the parameter space of the electronic
momentum k,

γj (k) = dk · Aj (k). (29)
C

The integral is gauge-invariant for a closed path, while the Berry vector potential,
or the Berry connection


Aj (k) = i uj k uj k ,
(30)
∂k

is a gauge-dependent quantity. This value, treated in analogy to electrodynamics as


a vector potential, gives access, using Stokes’ theorem, to a gauge-invariant quantity
called Berry curvature playing the role of a magnetic field in the parameter space:

Ω j (k) = ∇ k × Aj (k), (31)

or, alternatively,
4 Electronic Structure: Metals and Insulators 205

Ωj,μν (k) = ∂μ Aj,ν − ∂ν Aj,μ = −2 ∂μ uj k | ∂ν uj k


, (32)

γj = dS · Ω j (k), (33)
S

with the integration over an arbitrary surface enclosed by the path C.


In case of a translation-invariant crystal, a closed path in Eq. (29) occurs due
to a torus topology of the Brillouin zone as any two points k and k + G are
fully equivalent for any reciprocal lattice vector G, leading to an integration over
a Brillouin zone in Eq. (29).
Concerning applications of the Berry phase concept, we focus here on the
electron dynamics in the presence of an electric field E entering the Hamiltonian
trough a time-dependent uniform vector potential A(t), preserving the translation
symmetry of the system. The group velocity of a state (j, k) is given to first
order by an expression [65, 68] consisting of two terms, the usual band dispersion
contribution as well as a so-called anomalous velocity proportional to the Berry
curvature of the bands,

∂j (k) e
v j (k) = − E × Ω j (k), (34)
h̄∂k h̄

with the Berry curvature Eq. (32) given in the form [69, 67]:

Ω j (k) = i ∇ k uj k | × |∇ k uj k
. (35)

The Berry curvature is nonzero either in non-centrosymmetric systems or in systems


with broken time-reversal symmetry and vanishes in the case when both symmetries
are present, leading to elimination of the anomalous velocity in Eq. (34) [67]. Thus,
due to these symmetry properties of the Berry curvature, the anomalous Hall effect
can be observed in ferromagnetic systems. One has to stress the crucial role of
the spin-orbit interaction required for a nonzero Berry curvature in FM systems,
leading to avoided crossings of the energy bands which give in turn the most
pronounced contributions in the vicinity of the Fermi surface that can be seen in
Fig. 6a [70].
As one can see in Eq. (34), the anomalous velocity is always transverse to the
electric field giving rise to a Hall current. The corresponding contribution to the
Hall conductivity associated with the anomalous velocity term was demonstrated
first by Karplus and Luttinger [72] and is called Karplus-Luttinger mechanism. In
terms of the Berry curvature, it is given by the expression [65, 68, 69, 73]:

e2 dDk  γ
KL
σαβ = f (j (k))αβγ Ωj (k), (36)
h̄ BZ (2π )D n

with D the dimensionality of the system and αβγ the Levi-Civita tensor. As one
can see, this contribution is determined by all occupied states, and it is the only term
206 H. Ebert et al.

Fig. 6 (a) Fermi surface in the (010) plane (solid lines) and integrated Berry curvature −Ω z (k)
in atomic units (color map) of fcc Fe. (From Yao et al. [70]); (b) Berry curvature projected onto
the (k3 , k1 ) plane for Mn3 Ge, where k3 and k1 are aligned with the kz and kx axes, respectively.
It is calculated by integrating along the k2 direction. (From Ref. [71]); (c) Energy dispersion of
Mn3 Ge along k2 with (k3 , k1 ) fixed at the point with the largest Berry curvature, indicated by a
black dashed circle in (b). (From Ref. [71]). (Figure are printed with permission from [70] by
the American Physical Society; Figure (b) and (c) reprinted with permission from [71] by the
American Physical Society.)

contributing to the AHC for insulating systems. It is independent on the nature of the
impurities and their concentration and therefore is called intrinsic contribution to the
AHE. Note, however, that this contribution is not the only intrinsic contribution in
metals. The so-called side-jump mechanism suggested by Berger [74] is contributed
by the electrons at the Fermi surface and also yields a Hall conductivity that is
independent of the impurity concentration [69].
Note that the anomalous thermoelectric transport driven by statistical forces due
to a temperature gradient ∇T (the same concerns also the transport driven by
chemical potential gradients ∇μ) cannot rely on the anomalous velocity term as
it vanishes in the absence of an electric field. A corresponding theory giving the
intrinsic contribution to transverse thermoelectric transport has been reported by
Xiao et al. [75]. It is based on the generalization to finite temperatures of a theory
giving a Berry-phase correction to the orbital magnetization [76, 77, 17]:
4 Electronic Structure: Metals and Insulators 207

 dDk
M(r) = f (j (k))mn (k)
BZ (2π )D
j

1 dDk e  
+ D
Ω j (k)log 1 + e−β(j (k)−μ) , (37)
β BZ (2π ) h̄
j

where mn (k) = (e/2h̄)i ∇ k uj,k |[j (k) − H(k)] × |∇ k uj,k


is the orbital moment
of state n. By doing some transformations, this expression can be written also in the
form [67]:

 dDk 1

Mz (r) = D
f (j (k))mn,z (k) + df ()σxy (). (38)
BZ (2π ) e
j

It has two different contributions associated with the self-rotation of the wave packet
representing an electron and with the center-of-mass motion, respectively. The first
term, obviously, occurs by treating the carrier as a wave packet having finite spread
in the phase space. The second term comes from the Berry-phase correction to the
electron density of states [67].
A crucial point for thermoelectric transport is that the conventional expression
used for the current density is incomplete as it is derived for the carrier treated as a
point particle. Having a corresponding expression for the local current J obtained by
treating the carrier as a wave packet, introducing the concept of a transport current

j = J − ∇ × M(r), (39)

and using the expression Eq. (37) for the orbital magnetization density M(r), one
obtains the transport current as given by:

dDk
j = −e D
g(r, k)ṙ (40)
BZ (2π )

1 dDk e
−∇ × Ω(k)log(1 + e−β((k)−μ) ). (41)
β BZ (2π )D h̄

With this expression, it is straightforward to calculate various thermoelectric


responses to statistical forces [67]. Thus, the expression for an anomalous Nernst
conductivity αxy is given in terms of the intrinsic anomalous Hall conductivity σxy :

1 ∂f −μ
αxy =− d σxy () . (42)
e ∂μ T

Talking about the AHE in antiferromagnets (AFM), one has to distinguish the
systems according to their symmetry. The AHE does not occur in collinear AFMs
symmetric with respect to time-reversal symmetry T combined with a half-magnetic
208 H. Ebert et al.

unit-cell translation Ta/2 or spatial inversion I, i.e., either Ta/2 T symmetry or


IT (see [78]). However, in AFM materials having a symmetry violating these
conditions, the AHE can be observed. Accordingly, a large intrinsic AHE was
predicted in Mn3 Ir [79] and Mn3 Ge [80] by performing first-principles electronic
structure calculations. As already indicated, the avoided band crossings near the
Fermi surface give the dominant contribution to the AHE. An example of a
calculated k-resolved Berry curvature for Mn3 Ge [71] is plotted in Fig. 6b. It is
dominating in the area highlighted in red giving rise to the large AHC. In order
to investigate the origin of the hot spot at (0.127, 0.428), the band structure is
plotted in Fig. 6c along k2 varying from 0 to 1. As one can see, the Fermi level
crosses two small gaps around k2 = 0 and 0.5. This implies that the entanglement
between occupied and unoccupied states must be very strong around these two
points, giving a large contribution to the Berry curvature and in turn to the AHE.
Finally, it should be noted that also the real-space topology of a material as, for
example, the presence of a noncoplanar, chiral spin texture can give rise to an
intrinsic AHC. This phenomenon, that requires neither a finite external magnetic
field, nor a finite net magnetic moment, nor even spin-orbit coupling, is commonly
termed topological or chirality-induced Hall effect. The newly emerging field of
topological antiferromagnetic spintronics [81] deals with this and related response
phenomena at the intersection of antiferromagnetic spintronics with topology.

Itinerant Magnetism of Solids

Historically there was for a long time a heated discussion whether the model of
itinerant or band magnetism is a suitable platform to discuss the magnetic properties
of a specific solid or whether the assumption of local magnetic moments is more
appropriate. With the advances in electronic structure theory to provide schemes
that allow treating the electronic structure of solids from wide-band metallic solids
up to narrow-band oxides, including localized systems showing the formation of
electronic multiplets, the competition between these extreme models gets more
or less obsolete. Accordingly, this section gives only a brief introduction to the
theory of itinerant or band magnetism on the basis of the Stoner model followed
by discussing the main electronic features of two prototype class of materials:
disordered alloys of transition metals and the Heusler alloys.

Stoner Model of Itinerant Magnetism

A rather simple criterion for the spontaneous formation of ferromagnetic order is


provided by the Stoner model for itinerant magnetism. Starting point is the spin-
projected DOS for a paramagnetic solid as sketched in Fig. 7. Application of an
external magnetic field Bext gives rise to the Zeeman splitting ΔEZ = 2μB Bext for
spin-up and spin-down states with μB the Bohr magneton. The flip of the spin for
some states in the vicinity of the Fermi energy EF reestablishes a common Fermi
4 Electronic Structure: Metals and Insulators 209

E E E

EF EF EF

2 μBBext
n↓(E) n↑(E) n↓(E) n↑(E) n↓(E) n↑(E)

Fig. 7 Left: spin-dependent DOS n(EF ) at the Fermi energy EF for a paramagnetic metal. Middle:
Zeeman splitting ΔEZ = 2 μB Bext due to an external magnetic field Bext . Right: spin flip leads to
a common Fermi energy EF and a finite spin-magnetic moment M = (N↑ − N↓ )

energy for both spin systems leading to a net spin-magnetic moment M = (N↑ −
N↓ ). For small magnetic fields Bext , the resulting Pauli spin susceptibility χ =
M μB /Bext is determined by the density of states n(EF ) at the Fermi energy EF :

χ0 = 2μB n(EF ). (43)

An imbalance of N↑ and N↓ also changes the total energy due to the exchange
interaction. The corresponding spin-dependent correction for the electron energies
Ej kσ for spin σ (with σ = ±1/2) may be written as [82]:

Ej kσ = Ej k + sign (σ ) M I (44)

with the Stoner exchange integral I originally seen as a parameter. Accounting for
this correction, in addition one is led to the enhanced spin susceptibility

χ = S χ0 , (45)

with S the Stoner enhancement factor:


1
S= . (46)
1 − I n(EF )

The paramagnetic state remains stable as long as I n(EF ) < 1 holds. However, when
I n(EF ) approaches the value 1, the enhancement factor S and with this the induced
magnetic moment diverge indicating an instability. In fact, I n(EF ) > 1 for the
paramagnetic reference state implies that the increase of kinetic energy associated
with the flip of the spin for electrons at the Fermi energy is more than compensated
by the resulting change in the exchange-correlation energy even without an external
field leading to a stabilization of the ferromagnetic state [83]. Accordingly, the
Stoner criterion I n(EF ) > 1 indicates the spontaneous formation of ferromagnetic
spin order for a solid.
210 H. Ebert et al.

Linear response theory allows deriving explicit expressions for the static [84,
85, 86] and dynamic [87, 88] spin susceptibility of arbitrary systems that confirm
Eqs. (43) and (45) even for inhomogeneous, i.e., non-bulk, systems. Working within
the framework of SDFT provides in particular a clear prescription for the calculation
of the Stoner exchange integral [89, 84]:

I= d 3 r γ (r)2 K(r), (47)

with the induced spin polarization γ (r) and the exchange-correlation kernel K(r):

γ (r) = |ψj k (r)|2 δ(EF − Ej k )/n(EF ) (48)
jk
 
1 δ 2 Exc
K(r) = − . (49)
2 δm2 m(r)=0

Corresponding numerical results for the Stoner exchange-correlation integral I


(left) and density of states n(EF ) (right) are given in Fig. 8 for the 3d, 4d, and 5d
transition metal rows. As one notes, I varies smoothly within a transition metal
row. Accordingly, the product I n(EF ) primarily reflects the variation of n(EF )
with atomic number. In line with experiment, the Stoner criterion for ferromagnetic
ordering is met only for the late 3d transition metals Fe, Co, and Ni. Because of
the increase of the d-band width when going from a 3d metal to the corresponding
isoelectronic 4d or 5d metal, the Stoner product decreases and with this the tendency
toward ferromagnetic ordering. This trend is best seen for the sequence of fcc-metals
Ni-Pd-Pt that leads from a ferromagnet to strongly enhanced Pauli paramagnets with
a Stoner enhancement factor of 5.96 and 2.16, respectively.

1
3d 2 3d
4d 4d
0.9 5d 5d
n(EF) (1/eV)

1.5
0.8
I (eV)

0.7 1
0.6
0.5
0.5

0.4 0
Ti V Cr Mn Fe Co Ni Cu Ti V Cr Mn Fe Co Ni Cu
Zr Nb Mo Tc Ru Rh Pd Ag Zr Nb Mo Tc Ru Rh Pd Ag
Hf Ta W Re Os Ir Pt Au Hf Ta W Re Os Ir Pt Au

Fig. 8 Stoner exchange-correlation integral I (left) and density of states n(EF ) (right) at the Fermi
energy EF for the 3d, 4d, and 5d transition metal rows [26]
4 Electronic Structure: Metals and Insulators 211

As the Stoner integral I in general does not change much with the atomic
environment, the Stoner factor is primarily determined by the DOS n(EF ) at the
Fermi energy EF . As the d-band width W of the transition metals decreases with
coordination number, this leads usually to an increase of n(EF ). This implies
a corresponding increase of the tendency toward ferromagnetic ordering with
reduced dimensionality. Accordingly, magnetically ordered surface layers have been
predicted by theory for the paramagnetic metals V and Pd while the experimental
situation is unclear. For free and deposited transition metal clusters, many SDFT-
based calculations led to finite spin-magnetic moments as, for example, for free
Ru13 , Rh13 , and Pd13 clusters [90]. Additional calculations for the paramagnetic
state gave a large peak for the DOS near the top of the valence band with the Fermi
energy located at its maximum, i.e., the magnetic ordering could be explained on
the basis of the Stoner criterion. These results are fully in line with experimental
data for RuN and PdN clusters; for example, mRh ≈ 0.8 μB /atom in Rh9 , mPd <
0.4 μB /atom in Pd13 and mRu < 0.32 μB /atom in Ru10 clusters [91]. In accordance
with the Stoner criterion, an increase of cluster size led to a decrease of magnetic
moments as, for example, mRh ≈ 0.16 μB /atom in Rh34 and mRu < 0.09 μB /atom
in Ru1115 clusters.
The Stoner criterion was also applied successfully to deposited atoms. For exam-
ple, in line with experiment, self-consistent LSDA-based calculations predicted a
finite and vanishing spin-magnetic moment for Fe and Ni atoms, respectively, on a
chalcogenide topological insulator surface [92].

Slater-Pauling Curve

The substitutional magnetic alloys of 3d transition metals are often seen as prototype
materials for itinerant metallic magnetism. The experimental data on the average
magnetic moment M per atom of these systems is summarized by the well-known
Slater-Pauling curve shown in Fig. 9 (top). There are two main branches to be seen:
one leading from Fe to Cu with a slope of −45◦ and another one from Fe to Cr
with a slope of +45◦ . M is obviously given by M = N↑ − N↓ , where Nσ is the
average number of valence electrons with spin character σ . With the total number
of valence electrons Z = N↑ + N↓ , one gets the simple relation M = 2 N↑ − Z. On
the basis of the outdated rigid-band model that postulates a common electronic band
structure for both components of a binary alloy, all Ni and Co alloys are considered
to be strong ferromagnets with their spin-up band filled. Accordingly, N↑ is constant
and M decreases with Z explaining the right main branch of the Slater-Pauling
curve. For Z = 8.25, one may assume that the Fermi energy is at the top of the
spin-up band. Decreasing Z, one can now expect that N↓ stays constant leading to
M = Z − 2 N↓ . This obviously gives a simple explanation for the second branch
including in particular the Fe-Cr and Fe-V alloys.
Using the tight-binding version of the CPA combined with the Hartree-Fock
approximation, Hasegawa and Kanamori [94] could already give an alternative qual-
itative explanation for the Slater-Pauling curve avoiding the unrealistic assumptions
212 H. Ebert et al.

2.5 FeCo
Experiment FeTi
FeNi
2
magnetic moment (µ B)

CoFe

1.5 NiFe NiCo


FeMn
CoCr
1
CoMn
FeCr NiMn
0.5 FeV NiTi NiCu

NiCr
0
NiV
24 25 26 27 28
electron number / atom

2.5
FeCo
Theory
2 FeNi
magnetic moment (µ B)

FeCu
NiFe CoFe
FeSc
1.5 CoMn(2)
NiCo
CoCr
1
FeTi CoMn (1) NiMn
0.5 NiTi NiCu
FeV FeCr
NiCr
0 NiV
24 25 26 27 28
electron number / atom

Fig. 9 Top: experimental Slater-Pauling curve, i.e., the average magnetic moment corresponding
to the saturation magnetization of Fe-, Ni-, and Co-based alloys vs. average number of electrons
per atom. Bottom: corresponding theoretical results obtained by means of the KKR-CPA. The fcc,
instead of hcp, structure is assumed for Co-based alloys. For Co-Mn, two solutions, CoMn(1) with
a Mn local moment parallel to the bulk magnetization and CoMn(2) with an antiparallel moment,
are obtained. (All data taken from [93])

of the rigid-band model. This approach could be improved a lot by Akai using
the KKR-CPA within the framework of SDFT leading for most alloy systems to
a quantitative agreement between theory and experiment [93] (see bottom panel of
Fig. 9). Most importantly, this implies that the average moments are well described
by the effective CPA medium.
As found by experiment, the curves on the left-hand side, connected with Fe-
based bcc alloys, have a slope of about +45◦ although there is some spread. The
common feature of the alloys belonging to these subbranches is that the solute
atoms (Sc, Ti, V, and Cr) have negative local magnetic moments; i.e., they are
aligned antiparallel to the moments of the host (Fe, Co, and Ni). These negative
4 Electronic Structure: Metals and Insulators 213

moments can be associated with the appearance of hole states above or at the top
of the majority-spin d-bands. These hole states have a large amplitude at the solute
atoms causing a negative local moment there, as well as the rapid decrease of the
average moment with the solute concentration. The existence of the hole states in
the majority-spin band also affects the concentration dependence of the average
moment. In the case that no holes exist in the majority-spin band, the main origin
of the concentration dependence of the average moment is the reduced number of
available d-electrons as determined by the average number of valence electrons. In
fact, the straight line with a slope of −45◦ in the right half of the Slater-Pauling
curve, that is mainly associated with Ni-based fcc alloys, is explained this way.
Contrarily, in the case that holes exist in the majority-spin band, it is mainly the
missing number of majority d-states that causes the concentration dependence.
Thus, the slope of the average moment against the total number of electrons varies
depending on the solute atoms; the smaller the difference in the number of the
valence electrons, the steeper the slope. The above discussion, however, is useful
only for simple cases where the magnetic state is rather stable, typically in the
region of the strong ferromagnetism of Ni. More delicate situations as, for example,
the Ni-Fe, Ni-Mn, and Fe-Mn alloys in the vicinity where ferromagnetism becomes
instable, however, need a more detailed and specific discussion [93].
An important feature of the CPA calculations is that the alloy components
essentially keep their intrinsic properties. Fe, Co, and Ni, for example, have in
general in the various alloys a spin moment close to that of the pure elements (2.2,
1.7, and 0.6 μB , respectively; see Table 1). This is fully in line with results of neutron
scattering experiments or XMCD (X-ray magnetic circular dichroism) experiments
that give access to the spin and orbital moment in an element-specific way via the
XMCD sum rules [9]. Figure 10 shows corresponding results for the average spin-
and orbital magnetic moment per atom in fcc-Fex Ni1−x together with component-
specific data as calculated via the relativistic KKR-CPA on the basis of the LSDA
and LSDA+DMFT, respectively, in comparison with experiment [95]. As one notes,
the individual spin-magnetic moments of Fe and Ni in fcc-Fex Ni1−x show only a
rather weak concentration dependence. This also applies for the spin-orbit-induced
orbital moments that are defined by the expectation value of the angular momentum
operator lz (see, e.g., the discussion in [17]). These findings are in full agreement
with the individual moments determined via XMCD at the L2,3 -edges of Fe and Ni
[95]. Here, it is interesting to note that the theoretical results for the spin moment
hardly depend on the computational mode, i.e., whether the calculations are based
on the LSDA or the LSDA+DMFT. The orbital moment, on the other hand, depends
strongly on the computational mode. In particular it is found that inclusion of
correlation effects via the DMFT improves agreement with experiment.

Heusler Alloys

There are plenty of experimental and theoretical investigations on Heusler alloys in


the literature because of their very rich variety of magnetic properties including
214 H. Ebert et al.

LSDA
2 0.12 LSDA+DMFT
Expt

1.5
mspin (μB)

morb (μB)
0.08
1
LSDA
LSDA+DMFT 0.04
0.5 Expt

0 0
0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8
xFe xFe
3
0.2 LSDA
Fe Fe LSDA+DMFT
Expt
2.5 0.15
mspin (μB)

morb (μB)

2 0.1

1.5 LSDA 0.05


LSDA+DMFT
Expt
1 0
0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8
xFe xFe
0.25
1
Ni LSDA
Ni LSDA+DMFT
0.2 Expt
0.8
mspin (μB)

morb (μB)

0.6 0.15

0.4 0.1
LSDA
0.2 LSDA+DMFT 0.05
Expt
0 0
0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8
xFe xFe

Fig. 10 Top row: average spin- (left) and orbital (right) magnetic moment per atom in fcc-
Fex Ni1−x as calculated on the basis of the LSDA and LSDA+DMFT, respectively, in comparison
with experiment. In addition, the individual moments calculated for Fe (middle row) and Ni (bot-
tom row) are given. The experimental component-specific data stem from XMCD measurements.
(All data taken from [95])

ferromagnetism, antiferromagnetism, helimagnetism, and Pauli paramagnetism.


Depending on their crystal structure, one can distinguish two families of Heusler
alloys: semi-Heuslers of the type XYZ with C1b structure and full-Heuslers of the
type X2 YZ with L21 structure. Most of the Heusler alloys are metals; however, for
some systems also, half-metallic and semiconducting behavior has been observed.
The half-metallicity found in certain Heusler magnets [62, 87, 96, 97, 98]
attracted especially strong interest over the last 30 years because of its possible
4 Electronic Structure: Metals and Insulators 215

Fig. 11 Left: spin-resolved DOS for Co2 MnSi showing a bandgap for the minority states. Right:
comparison of the spin polarization obtained by in situ SRUPS on a Co2 MnSi thin film with
the calculated DOS-derived spin polarization, the calculated UPS spin polarization including
broadening effects and considering only bulk states, and the calculated total UPS spin polarization
including broadening effects with additional surface state contributions [99]

use in spintronics and magneto-electronics. Half-metallic materials exhibit metallic


behavior only for one spin direction, while for the other spin direction the Fermi
level is located in a bandgap. Figure 11 shows the spin-resolved DOS for Co2 MnSi
as a representative example. Defining the spin polarization p of a material in terms
of the spin-dependent density of states nσ (E) according to

n↑ (E) − n↓ (E)
p= (50)
n↑ (E) + n↓ (E)

one has for half-metallic materials 100% spin polarization at the Fermi energy that
should lead to a fully spin-polarized electric current.
For real materials, however, the spin polarization may be influenced in various
ways. Mavropoulos et al. [100], for example, demonstrated the impact of spin
mixing caused by spin-orbit coupling for the Heusler alloys of the type XMnSb.
As expected, this influence increased with increasing atomic number of the element
X, as reflected by the ratio n↓ (EF )/n↑ (EF ) = 0.25, 0.30, 0.35, 0.75, and 2.70 found
for the series X = Co, Fe, Ni, Pd, and Pt. As a consequence, one can expect a spin
polarization well below 100% for compounds with heavier elements.
The influence of the surface on the spin polarization has been studied by
Galanakis [101] investigating the (001) surfaces of the semi-Heusler alloys
NiMnSb, CoMnSb, and PtMnSb and for the full-Heusler alloys Co2 MnGe,
Co2 MnSi, and Co2 CrAl. In general, a rather strong modification of electronic
and magnetic properties has been found for the surface region. In the case of semi-
Heuslers, the Ni-, Co-, or Pt-terminated surface has a rather large DOS at the Fermi
level for minority-spin states, while for the MnSb-terminated surfaces the calculated
properties are close to those of the bulk. Nevertheless, half-metallicity disappears
also in this case due to surface states, resulting in a spin polarization at the Fermi
216 H. Ebert et al.

level of 38%, 46%, and 46% for NiMnSb, CoMnSb, and PtMnSb, respectively. A
similar behavior was found for the Co-terminated surfaces of the full-Heusler alloys
Co2 MnGe, Co2 MnSi, and Co2 CrAl. In the case of the MnGe-terminated (001)-
surface of Co2 MnGe, the spin polarization vanishes also due to the surface states,
although the CrAl termination of Co2 CrAl leads to a very high spin polarization of
around 84%.
In experiment, the spin polarization of a material can be determined among
others by tunneling experiments or by spin-resolved photoemission [102, 103, 104].
Figure 11 shows corresponding experimental and theoretical spin-polarization
curves of Co2 MnSi [99] for the UPS-regime (hν = 21.2 eV) that showed for the
surface regime a value of 93% at the Fermi level, the largest observed so far.
NiMnSb was among the first semi-Heusler compounds predicted to be half-
metallic [105] and was intensively investigated since then by experiment as well
as theory [106, 107, 108]. The Sb atoms with the atomic configuration 5s2 5p3 lead
to the formation of a deep lying narrow s- and a p-band at around 12 and 3 − 5 eV,
respectively, below the Fermi energy EF . Accordingly, these states are not involved
in the formation of the bandgap near EF . The Ni and Mn d-states hybridize with
each other as well as with the sp-states of Sb leading to the formation of bonding
and antibonding bands. For the paramagnetic state of NiMnSb, the Fermi level
lies in the middle of an antibonding band mainly associated with the d-states of
Mn. Accordingly, exchange splitting of the antibonding band leads to a gain in
energy accompanied by formation of a strong magnetic moment for the Mn atom.
As a result, the Fermi energy moves to the energy gap separating bonding and
antibonding minority-spin states. Due to this, there are nine minority-spin states per
unit cell below EF , one due to the Sb-like s-band, three due to the Sb-like p-bands,
and five due to the Ni-like d-bands, that are all occupied. As atoms forming the alloy
contribute 22 electrons per unit cell, the majority-spin band contains 22 − 9 = 13
electrons, resulting in a moment of 4 μB per unit cell.
Plotting the total spin-magnetic moment per unit cell, M, of NiMnSb together
with that of other semi-Heusler half-metallic compounds as a function of the total
number of valence electrons Z, one can see that M – in analogy to the Slater-Pauling
curve for the binary transition metal alloys – follows the relation: M = Z −18 [107]
(see Fig. 12 (left)). This relation is a consequence of the complete occupation of the
nine minority-spin bands and follows directly from the definitions Z = N↑ + N↓
and M = N↑ − N↓ that lead to mt = Z − N↓ [98].
The occurrence of half-metallicity in the case of the prototype full-Heusler
compounds Co2 MnSi and Co2 MnGe was also predicted by electronic structure
calculations [109, 110]. Similar to the case of semi-Heusler alloys, the sp-bands
of Si and Ge are located well below EF . For that reason, they do not participate in
the formation of the energy gap that is caused by the hybridization of Mn and Co d-
states. As a result, for the spin-polarized state of the material, the Fermi level is again
located within the minority-spin energy gap, so that the minority band contains 1 s-
band and 3 p-bands derived from the sp-element and 8 Co-related d-bands, which
are fully occupied by 12 electrons [98]. As shown by Galanakis et al. [111], the
spin-magnetic moment of the full-Heusler alloys accordingly follows the relation
M = Z − 24. Corresponding theoretical and experimental data are summarized in
4 Electronic Structure: Metals and Insulators 217

6 6
Co2MnSi
Co2MnAs
Co2MnGe Co2MnSb
Co2CrAl Co2MnSn Co2FeSi
NiMnSe
NiMnSb Fe2MnSi Co2FeAl
NiMnTe Ru2MnSi Rh2MnIn
PdMnSb
4 PtMnSb
4 Ru2MnGe Rh2MnTl
Ni2MnAl
CoFeSb Ru2MnSn
RhMnSb NiFeSb Rh2MnGe
Co2VAl Rh2MnSn
M

M
CoMnSb 2 Fe2MnAl Rh2MnPb
IrMnSb Co2TiSn
2 FeMnSb
NiCrSb Fe2CrAl Co2MnAl
CoCrSb Co2TiAl Co2MnGa
0

24
CoVSb NiVSb Fe2VAl Rh2MnAl

Z-
18

Mn2VGe Rh2MnGa

=
Z-

Ru2MnSb

M
0 CoTiSb
=

-2 Mn2VAl
M

16 18 20 22 24 20 22 24 26 28 30 32
Z Z

Fig. 12 Left: calculated total spin moments per unit cell for several semi-Heusler alloys.
Experimental values are given for NiMnSb (3.85 μB ), PdMnSb (3.95 μB ), PtMnSb (4.14 μB ), and
CoTiSb (nonmagnetic) [107]. Right: calculated total spin moments for several full-Heusler alloys.
Experimental values are given for Co2 MnAl (4.01 μB ), Co2 MnSi (5.07 μB ), Co2 MnGa (4.05 μB ),
Co2 MnGe (5.11 μB ), Co2 MnSn (5.08 μB ), Co2 FeSi (5.9 μB ), Mn2 VAl (1.82 μB ), and Fe2 VAl
(nonmagnetic) [111]. In both figures the dashed line represents the corresponding Slater-Pauling
curve. The open circles represent the compounds deviating from this curve

the Slater-Pauling plot in Fig. 12 (right). The difference to the Slater-Pauling curve
of the binary transition metal alloys in Fig. 9 is due to the fixed number of minority-
spin electrons in the half-metallic Heusler compounds. In this case, increasing Z
leads to a filling of the majority band, while for the ferromagnetic transition metal
alloys, the relation M = 10 − Z for the right branch is a result of the full occupation
of five majority-spin d-states and charge neutrality achieved by filling the minority-
spin d-states.
Within the family of Heusler compounds, there are furthermore the so-called
inverse full-Heusler compounds that have a similar chemical formula X2 YZ but
crystallize in the so-called Xα structure [98]. The prototype of the inverse full-
Heusler compounds is Hg2 TiCu [112]. As Fig. 13 demonstrates, the magnetic
moments per unit cell as a function of valence electrons also follow in this case
corresponding Slater-Pauling rules.

Total Electronic Energy and Magnetic Configuration

The calculation of the electronic total energy allows to seek for the magnetic ground
state of a solid. This applies to the crystal structure as well as to the specific magnetic
ordering or spin configuration, respectively. For many purposes, it is helpful to
represent the neighborhood of the ground state or a suitable magnetic reference
state by mapping the complex configuration dependence of the total energy of the
system on an approximate spin Hamiltonian. An issue in this context is spin-orbit
coupling that removes energetic degeneracies of competing spin configurations, may
218 H. Ebert et al.

Mn2NiAs
M = Z-18 M = Z-24 Mn2CoAs
4 Ti2NiAl Ti2CuAl
4 Mn2NiSi
Cr2CoAs
Ti2CoAl Ti2CoSi Mn2FeSi Mn2FeAs
Sc2NiAl Ti2FeSi Ti2FeAs Mn2CoSi
Mn2MnAs
2 Sc2CoSi V2MnAl V2CrSi 2 Cr2NiSi
Ti2FeAl Mn2CoAl
Sc2NiSi V2CoSi
V2CrAl V2FeAs Cr2NiAl
Ti2MnAs V2NiAs Cr2CoSi
Cr2CrAs
M

M
0 Ti2CrAl Ti2MnAl 0 Cr2FeAl Cr2CoAl Cr2FeAs
Sc2MnAl Ti2VAs Cr2MnAs Mn2FeAl
Sc2CrSi Ti TiSi Sc2FeAl Mn2CrAl V2NiAl Mn2MnSi
V2VAl Mn MnAl
Sc2VAs 2 Sc2MnSi Cr2MnSi Mn2CrAs
Cr2CrSi 2
Mn2CrSi
-2 Ti2VAl Sc 2
CrAs -2 V2MnSi V2CoAl
Ti2VSi V2FeSi
Sc2VSi V2FeAl
Sc2VAl Sc2CrAl V2MnAs
Cr2MnAl
-4 -4 Cr2CrAl

13 15 17 19 21 23 20 22 24 26 28 30
Z Z

Fig. 13 Total spin-magnetic moments per unit cell (in μB ) as a function of the total number
of valence electrons Z in the unit cell for several compounds. The lines represent the two of the
various forms of the Slater-Pauling rule [98]. The compounds within the frames follow one of these
rules and are perfect half-metals, while the rest of the alloys slightly deviate. For this reason, their
total spin-magnetic moment is represented by an open red circle. The sign of the spin-magnetic
moments has been chosen so that the half-metallic gap is in the spin-down band

lead to the anisotropic Dzyaloshinsky-Moriya exchange interaction, and gives rise


to magnetocrystalline anisotropy.

Total Electronic Energy and Magnetic Ground State

Access to the electronic total energy Etot (see Eq. (10)) in principle allows to
determine the magnetic ground state of any solid. As a corresponding example
for this, results of LSDA-based calculations for Fe in the para- (PM), ferro- (FM),
as well as antiferromagnetic (AFM) state with bcc and fcc structure are shown in
Fig. 14 [27]. Obviously, the use of the LSDA led to an over-binding, i.e., to a lattice
parameter that is too small and a bulk modulus that is too large when compared
with experiment (see Table 1). Most importantly, however, the paramagnetic fcc-
phase was found as the ground state instead of the ferromagnetic bcc-phase. Use of
the GGA on the other hand improved the situation very much giving in particular
the ferromagnetic bcc-phase as the ground state.
Another example for a search for the magnetic ground state configuration is given
by Fig. 15. In this case, the so-called fixed spin moment method was used to explore
the dependency of the total energy Etot on the lattice parameter and average atomic
moment of disordered fcc-Fex Ni1−x alloys. As one notes, a double minimum occurs
in the vicinity of the concentration x = 0.65 indicating the competition between a
low-volume, low-spin moment phase and a high-volume, high-spin moment phase.
In fact, this has been seen as a possible explanation for the occurrence of the invar
effect for that composition. Another study on the invar effect, however, stressed the
importance of a noncollinear spin configuration [114].
4 Electronic Structure: Metals and Insulators 219

400 600

PM bcc
PM bcc
200 400
E (meV)

E (meV)
FM bcc
0 200 PM fcc
PM fcc AFM fcc

FM bcc
-200 0
60 65 70 75 80 70 75 80 85 90
Volume (a.u.) Volume (a.u.)

Fig. 14 Total energy Etot of paramagnetic (PM) bcc and fcc, ferromagnetic (FM) bcc, and
antiferromagnetic (AFM) fcc-Fe as a function of the volume as obtained within the LSDA (left)
and GGA (right), respectively. The curves are shifted in energy so that the minima of the FM-bcc
curves, corresponding to Etot = 0, coincide. (All data taken from [27])

2.5 2.5

2.0 2.0
M avg ( μ Β/atom )

M avg ( μ Β/atom )

1.5 1.5

1.0 1.0

0.5 0.5
Fe 60 Ni 40 Fe 65 Ni 35
0.0 0.0
6.3 6.4 6.5 6.6 6.7 6.8 6.9 6.3 6.4 6.5 6.6 6.7 6.8 6.9
a (a.u.) a (a.u.)

Fig. 15 Total energy or binding surfaces for the disordered ferromagnetic alloys Fe60 Ni40 and
Fe65 Ni35 . (All data taken from [113])

The occurrence of a noncollinear spin configuration is quite common for actinide


compounds but also for a number of transition metal-based systems [115]. A
prominent example for this is Mn3 Sn with its hexagonal unit cell given in Fig. 16.
Scalar relativistic calculations gave for all shown noncollinear spin configurations
a total energy well below that of competing collinear spin configurations. Ignoring
spin-orbit coupling, only the angle between the spin moments is relevant leading
to a degeneracy for all four spin configurations. However, if spin-orbit coupling is
taken into account, the degeneracy is removed and the resulting moment may deviate
to some extent from the orientation found for the scalar relativistic calculations
(see thin arrows for configuration (c) and (d) in Fig. 16). The presence of spin-
orbit coupling implies that the spin-magnetic moment is accompanied by an orbital
one. It is interesting to note that the orientation of the spin-orbit-induced orbital
magnetic moment may and will deviate from that of the spin moment if symmetry of
the system allows. Another interesting property of spin-compensated noncollinear
220 H. Ebert et al.

Mn Sn
z=1/4
z=3/4

(a) (b) (c) (d)

1 1 1 1

2 3 2 2 2
3 3 3

Fig. 16 Crystal and magnetic structure of Mn3 Sn. Rotations of the magnetic moments leading
to weak ferromagnetism in the structure in (c) and (d) are shown only for atoms in the z = 0.25
plane (thin arrows). Moments of the atoms in the z = 0.75 plane are parallel to the moments of the
corresponding atoms of the z = 0.25 plane. (All data taken from [115])

antiferromagnets having certain magnetic space groups is the occurrence of the


anomalous Hall effect that is usually not expected for an antiferromagnet [79].
However, group theoretical considerations unambiguously show that the Hall effect
may even show up for spin-compensated solids [116]. The first observations of the
AHE [117] as well as of its thermoelectric analog, the anomalous Nernst effect
[118], in a non-collinear antiferromagnet could in fact both be made in Mn3 Sn. The
occurrence of spin-polarized currents in this and related materials currently attracts a
lot of attention as well [119,120]. Finally, a criterion for the instability of a collinear
spin structure with respect to a transition to a noncollinear one was formulated by
Sandratskii and Kübler: If the collinear magnetic structure under consideration is
not distinguished by symmetry compared with the noncollinear structures obtained
with infinitesimal deviations of the magnetic moments from collinear directions, this
structure is unstable [121].

Exchange Coupling Parameters

When dealing with competing magnetic configurations, it is not always possible or


not necessary to perform full ab-initio calculations. In these cases, one may adopt a
multi-scale approach that uses the classical Heisenberg Hamiltonian:
4 Electronic Structure: Metals and Insulators 221


H=− Jij m̂i · m̂j , (51)
i=j

with m̂i(j ) the orientation of the magnetic moment on the lattice site i(j ), for
corresponding simulations. The isotropic exchange coupling parameters Jij , on
the other hand, are calculated in an ab initio way. This can be done by applying
a corresponding version of the so-called Connolly-Williams method [122]. This
implies to calculate the total energy for many different magnetic configurations
within a super-cell and to determine the exchange coupling parameters by fitting
the energy on the basis of Eq. (51). This way one obviously achieves a mapping
of the complicated energy landscape E({m̂i }) on the rather simple expression in
Eq. (51) involving only pair interactions that is easy to handle. Accordingly, a more
accurate representation of the energy as a function of the magnetic configuration
can therefore be expected by a cluster expansion as suggested by various authors
[123, 124].
Another approach to determine the exchange coupling parameters Jij in Eq. (51)
is to consider the change of the single-particle energy ΔEij if two magnetic
moments on sites i and j change their relative orientation. The necessary formal
developments started with the work of Oguchi et al. who expressed the difference
in energy between the ferro- and antiferromagnetic state of a solid making use of
multiple scattering theory and Lloyd’s formula [125]. Lichtenstein et al. [126]
extended this approach dealing with the coupling energy ΔEij associated with an
individual pair (i, j ) of atoms. If ΔEij is expressed to lowest order with respect to
the orientation angle of the moments m̂i and m̂j , one gets a one-to-one mapping of
the exchange coupling energy ΔEij to the Heisenberg Hamiltonian in Eq. (51), with
the exchange coupling constants Jij given by [126]:

 EF    
1 −1 −1
τ↑ tj−1 −1
ij ji
Jij =  dE Trace ti↑ − ti↓ ↑ − tj ↓ τ↓ , (52)

ij
where ti↑(↓) is the spin-dependent single site scattering matrix for site i and τ↑(↓) is
the spin-dependent scattering path operator matrix connecting sites i and j . Results
for the isotropic exchange coupling parameter Jij of Fe and Co as a function of
the distance Rij between sites i and j that have been obtained using an analogous
expression derived within the LMTO-GF formalism [127] are shown in Fig. 17. The
big advantage of this approach is that it can be applied with comparable effort to
more complex systems like disordered substitutional alloys [128], Heusler alloys
[129, 130, 131], diluted magnetic semiconductors [132, 133, 134, 135, 136, 137],
magnetic surface films [127,138], or finite deposited clusters [139,140]. In addition,
it should be mentioned that an approach similar to that leading to Eq. (52) was
worked out by Katsnelson and Lichtenstein [141] that allows for an improved
treatment of correlated systems.
If spin-orbit coupling is accounted for, the exchange coupling parameter in the
Heisenberg Hamiltonian has to be replaced by a corresponding tensor:
222 H. Ebert et al.

20 20

16 Fe 16 Co
12 12
Jij (meV)

Jij (meV)
8 8

4 4

0 0

-4 -4
1 1.5 2 2.5 1 1.5 2 2.5
Rij (units of lattice parameter) Rij (units of lattice parameter)

Fig. 17 Isotropic exchange coupling constants Jij of Fe and Co as a function of the distance Rij
between sites i and j . (All data taken from [127])

 
H=− m̂i J m̂j + K(m̂i ) (53)
ij
i=j i
 
=− Jij m̂i · m̂j − m̂i J S m̂j
ij
i=j i=j
   
− D ij · m̂i × m̂j + Ki (m̂i ), (54)
i=j i

with the accompanying single-site magnetic anisotropy represented by the term


K(m̂i ). In Eq. (53), the coupling tensor J has been decomposed into its isotropic
ij
part Jij , its traceless symmetric part J S , and its antisymmetric part. The latter
ij
one, that may occur in case of systems without inversion symmetry, is often
represented in terms of the so-called Dzyaloshinsky-Moriya (DM) vector D ij , with
βγ γβ
Dijα = 12 (Jij − Jij ) and a cyclic sequence of the Cartesian indices α, β, and
γ . A corresponding generalization of the nonrelativistic expression for Jij given in
Eq. (52) to its relativistic tensor form was worked out by various authors [18, 142]
and applied in particular to cluster systems [124, 143] with the interest focusing on
the impact of the DM interaction.
It should be emphasized once more that Eq. (51) and extensions to it supply an
approximate mapping of the complicated energy landscape E({m̂i }) of a system
calculated in an ab initio way onto a simplified analytical expression. This implies
corresponding limitations [124] in particular due to the use of the rigid spin
approximation (RSA) [144]. It is interesting to note that a coupling tensor of the
same shape as in Eq. (53) occurs for the indirect coupling of nuclear spins mediated
by conduction electrons. In this case the mentioned restrictions do not apply. As a
consequence, the linear response formalism on the basis of the Dyson equation can
4 Electronic Structure: Metals and Insulators 223

be used without restrictions to determine the corresponding nuclear spin-nuclear


spin coupling tensor [145].
Another approach using the same idea as the Connolly-Williams method is based
on the total energy ΔE(q, θ ) = E(q, θ ) − E(0, θ ) calculated for noncollinear spin
spirals (see section “Spin Spiral Calculations”) that are characterized by the wave
vector q and tilt angles (θ, φ(R i ) = q · R i ) for the moment mi on the atomic
position R i . In the case of a small tilt angle θ , ΔE(q, θ ) can be represented in terms
of the Fourier transform J (q) of the real space exchange coupling parameters [146]:

θ2
E(q, θ ) = E0 (θ ) − J (q). (55)
2
Accordingly, performing an inverse Fourier transformation, one can determine the
real-space interatomic exchange coupling parameters J0j . In the case of a simple
Bravais lattice, this is given by the expression

1  −iqR 0j
J0j = e J (q). (56)
N q

Uhl et al. [147] applied this scheme among others for a study of the invar system
Fe3 Pt. From their numerical results for ΔE(q, θ ), they evaluated the spin-wave
stiffness constant A and the exchange parameter J0 that allows to give an estimate
for the Curie temperature on the basis of the mean field approximation (MFA) (see
section “Finite-Temperature Magnetism”). Similar work was also done for two-
dimensional systems as for example magnetic surface films. Within their study on
an Fe film on W(110), Heide et al. also accounted for the spin-orbit coupling [148].
This way they could determine not only the spin-wave stiffness constant A but also
the Dzyaloshinsky-Moriya interaction vectors D.
As discussed in section “Spin Density Functional Theory”, dealing with systems
with narrow electronic energy bands, in order to go beyond the local spin-density
approximation, the LDA+U or +DMFT methods can be used to properly account
for strong electronic correlation in these materials. In order to adapt the method for
the treatment of exchange interactions formulated within the LSDA [126, 18, 142],
a corresponding theory was developed by Katsnelson and Lichtenstein [141] that
employs an analog of the local force theorem to derive expressions for effective
exchange parameters, Dzyaloshinsky-Moriya interaction, and magnetic anisotropy
in highly correlated systems. The authors demonstrated for the particular case of
ferromagnetic Fe that treating correlation effects beyond the LSDA (within the
LDA+Σ approach) in the exchange interactions results in a spin-wave spectrum
and spin-wave stiffness which are in better agreement with experiment than those
obtained within plain LSDA. The important role of additional contributions to the
exchange coupling of the correlation interactions has been demonstrated, e.g., for
half-metallic ferromagnetism in CrO2 [149], magnetic properties of CaMnO3 [150],
and magnetic properties of transition metal oxides [151]. Another important
feature of the exchange interactions observed in various materials are their strong
224 H. Ebert et al.

orientation dependence that would require to go beyond the Heisenberg model when
considering the finite-temperature or spin-wave properties of these materials. This
problem was recently discussed by different groups [151, 152, 153, 154].

Magneto-Crystalline Anisotropy

Magnetocrystalline anisotropy denotes the dependence of the total energy of


a system on the orientation m̂ of its magnetization with the anisotropic part
EA (m̂) of the energy taking a minimum for m̂ along a so-called easy direction
of the magnetization. Usually, EA (m̂) is split into the intrinsic material-specific
magnetocrystalline anisotropy (MCA) energy EMCA (m̂) and the extrinsic shape
anisotropy energy Eshape (m̂) determined by the shape of the sample:

EA (m̂) = EMCA (m̂) + Eshape (m̂). (57)

Considering the difference in energy ΔEX (m̂, m̂ ), with X=A, MCA or shape,
respectively, for the magnetization oriented along directions m̂ and m̂ , respectively,
one has accordingly ΔEX (m̂, m̂ ) = EX (m̂) − EX (m̂ ).
A convenient phenomenological representation of the magnetocrystalline
anisotropy energy can be given by an expansion in terms of spherical harmonics
Ylm (m̂)
 m=l

EMCA (m̂) = κlm Ylm (m̂) (58)
l even m=−l

or alternatively by an expansion in powers of the direction cosines (α1 , α2 , α3 ) =


(m̂ · x̂, m̂ · ŷ, m̂ · ẑ):
 
EMCA (m̂) = b0 + bij αi αj + bij kl αi αj αk αl + . . . (59)
i,j i,j,k,l

Assuming degeneracy of the energy upon time-reversal, i.e., flip of the magnetiza-
tion, only terms that are even with respect to the orientation m̂ of the magnetization
can occur in these equations. Further restrictions on the expansions are imposed by
the crystal symmetry of the investigated material [7]. Considering, for example,
a hexagonal system with the expansion up to sixth order, the corresponding
hex (m̂) are given by:
expressions for EMCA

hex
EMCA (m̂) = K̃0 + K̃1 Y20 (θ, φ) + K̃2 Y40 (θ, φ)
+K̃3 Y60 (θ, φ) + K̃4 Y64 (θ, φ) (60)
= K0 + K1 (α12 + α22 ) + K2 (α12 + α22 )2 + K3 (α12 + α22 )3
+K4 (α12 − α22 ) (α14 − 14α12 α22 + α24 ), (61)
4 Electronic Structure: Metals and Insulators 225

where the coefficients are interconnected by the relations K̃1 = 21


2
(7K1 + 8K2 +
8K3 ), K̃2 = 385 (11K2 + 18K3 ), K̃3 = 231 K3 and K̃4 = 10,395 K4 .
8 16 1

The shape anisotropy energy Eshape (m̂) is usually associated with the classical
dipole-dipole interaction of the individual magnetic moments mν on the lattice sites
[6, 155]. Accordingly, it can be determined straightforwardly by a corresponding
lattice summation:

 mν mν   
1 [R nνν  · m̂]2
Edip (m̂) = 1−3 , (62)
c2 |R nνν  |3 |R nνν  |2
νν  Rn

with R nνν  = R n + ρ ν − ρ ν  , where a periodic system has been considered with


lattice vectors R n and ρ ν the basis vectors within the unit cell.
For transition metal systems, the intrinsic part of the magnetic anisotropy
energy EMCA (m̂) has to be ascribed to the spin-orbit coupling. Accordingly, the
corresponding energy difference ΔESOC (m̂, m̂ ) can be determined by total energy
calculations with the magnetization oriented along directions m̂ and m̂ , respec-
tively, and taking the difference. Obviously, this implies a full SCF calculation for
both orientations and taking the difference of large numbers to get ΔESOC (m̂, m̂ ).
The problems connected with this approach can be avoided by making use of the
so-called magnetic force theorem that allows to approximate ΔESOC (m̂, m̂ ) by
the difference of the single particle or band energies (see Eq. (10)) for the two
orientations obtained using a frozen spin-dependent potential [6, 156]:

 EFm̂


ΔESOC (m̂, m̂ ) = − dE N m̂ (E) − N m̂ (E)

1    
− nm̂ (EFm̂ ) (EFm̂ − EFm̂ )2 + O(EFm̂ − EFm̂ )3 . (63)
2

Here EFm̂ is the Fermi energy for the magnetization along m̂ while nm̂ (E) and
E
N m̂ (E) = dE  nm̂ (E  ) are the corresponding DOS and integrated DOS,
respectively.
This approach is used extensively for compounds and layered systems and leads
typically to anisotropy energies that deviate less than 10% from results obtained
from full SCF calculations. By using, in the case of layered systems, layer-resolved
data for the DOS in Eq. (63), a corresponding layer decomposition of the anisotropy
energy ΔESOC (m̂, m̂ ) could be achieved [157]. Application of this scheme for the
spatial decomposition of ΔESOC (m̂, m̂ ) shows that the dominating contributions
originate in general from the interface or surface layers, respectively.
Equation (63) implies that spin-orbit coupling is accounted for within the
underlying electronic structure calculations. Instead one can start from a scalar
relativistic calculation and treat HSOC as a perturbation. Solovyev et al. [158] used
this approach on the basis of the Green function method in combination with the
Dyson equation Eq. (21). This allowed to write the spin-orbit-induced correction
226 H. Ebert et al.

ESOC (m̂) to the single-particle energies as a sum of two-site interactions:


 EF 
ESOC (m̂) = dE δN(E) = Eij (m̂) (64)
ij

with
 EF
1 ij j ji
Eij (m̂) = −  dE Trace G0 (m̂) HSOC G0 (m̂) HSOC
i
, (65)

ij
where G0 (m̂) are real space structural Green function matrices corresponding to the
scattering path operator in Eq. (20) [42]. In contrast to Eq. (63), Eq. (65) provides
a unique spatial or component-wise decomposition of the magnetocrystalline
energy. Solovyev et al. [158] used this approach for a detailed study of the
ordered compounds TX with T = Fe, Co and X = Pd, Pt having CuAu structure.
This way they could in particular show that the hybridization between the T
and X sublattices essentially determines their magnetocrystalline anisotropy. In
addition, an expression analogous to Eq. (65) allowed to demonstrate and discuss
the interconnection between the energy correction E(m̂) and the spin-orbit-induced
orbital magnetic moment μorb represented by the expectation value of the angular
momentum operator l. Using a similar approach as sketched here, this relation
was already investigated before by Bruno [159] and also by van der Laan [160].
Assuming a strong ferromagnet with the majority band filled, the relation:

1
ESOC (m̂) = − C ζ σ · l
, (66)
4

was derived, where C is a constant and ζ represents the strength of the spin-orbit
coupling. This equation was used in numerous experimental studies that exploited
the XMCD (X-ray magnetic circular dichroism) and the associated sum rules [9]
to determine in an element specific way the change of the angular momentum Δl
when changing the orientation of the magnetization from m̂ to m̂ to get a component
resolved estimate for the corresponding anisotropy energy ΔESOC (m̂, m̂ ).
A further approach to calculate the spin-orbit-induced anisotropy energy is to
consider the torque T (θ ) exerted on a magnetic moment m when the magnetization
is tilted by the angle θ away from its equilibrium orientation (easy axis). The
corresponding expression for T (θ ),

 ∂HSOC

T (θ ) =
ψj k ψj k , (67)
∂θ
j k occ

was given first by Wang et al. [161] for the case that the electronic structure is
represented in terms of Bloch states. A more general expression was obtained on
the basis of multiple scattering theory [162]:
4 Electronic Structure: Metals and Insulators 227


1 EF ∂

−1

Tαm̂û = −  dE ln det t( m̂) − G0
, (68)
π ∂α û

where the torque component with respect to a rotation of the magnetization around
an axis û is considered and where t(m̂) is the single-site t-matrix for an orientation
of the moments along m̂ and G0 is the corresponding free electron Green function
matrix.
On the basis of Eq. (67) or (68), respectively, the anisotropy energy
ΔESOC (m̂, m̂ ) is obtained from the torque by integrating along a path connecting
m̂ and m̂ . This approach is especially suited when dealing with systems with
uniaxial anisotropy. Neglecting in this case the dependence on φ, the anisotropy
energy can be represented as ESOC (θ ) = K0 + K1 sin2 (θ ) + K2 sin4 (θ ) with the
torque given by:

dESOC (θ )
T (θ ) = = K1 sin(2θ ) + 2K2 sin(2θ ) sin2 θ. (69)

For the special setting θ = π/4 and φ = 0, one has therefore:

ESOC (π/2) − ESOC (0) = K1 + K2 = T (π/4). (70)

This implies that if the contribution K1 sin2 (θ ) to ESOC (θ ) dominates, a situation


often met, K1 and with this ESOC (θ ) can be obtained from a single calculation for
the special settings. Otherwise, K1 and K2 can be obtained by a fit to a sequence of
calculations for varying angles θ .
The contribution ΔEdip (m̂, m̂ ) to the total anisotropy energy that is associated
with the dipole-dipole interaction of the individual magnetic moments is usually
treated classically by evaluating a corresponding Madelung sum (see Eq. (62)) [155,
163, 164]. While for most cases ΔESOC (m̂, m̂ ) is much larger than ΔEdip (m̂, m̂ ),
both contributions are often found for layered systems to be in the same order
of magnitude. As ΔEdip (m̂, m̂ ) always favors an in-plane orientation of the
magnetization while ΔESOC (m̂, m̂ ) in general favors an out-of-plane orientation,
one may have a flip of the easy axis from out-of-plane to in-plane with increasing
thickness of the magnetic layers. Such a behavior has been found, for example, for
Con Pdm multilayers as shown in Fig. 18 [163]. Similar results were obtained for the
magnetic surface layer system Fen /Au(001) that shows a change from out-of-plane
to in-plane anisotropy if the number n of Fe layers is larger than 3 [155].
In particular in cases for which ΔESOC (m̂, m̂ ) and ΔEdip (m̂, m̂ ) are of the
same order of magnitude, it seems questionable to treat the first contribution
quantum mechanically and the second one in a classical way. Although it was
pointed out already nearly 30 years ago that ΔEdip (m̂, m̂ ) is caused by the Breit
interaction [14], there is only little numerical work done in this direction [165,166].
Including a vector potential in the Dirac equation Eq. (23) that represents the
corresponding current-current interaction such numerical work has been done on
magnetic surface films and multilayer systems. It turned out in all investigated
228 H. Ebert et al.

Co1Pd2
2 0.68
Co1Pd5 0.4

ΔE (meV/unit cell)
Kt (mJ/m )

Co2Pd4

ΔE (meV)
2

1 0.34 0
Co4Pd2
-0.4
0 0
-0.8
Co3Pd3
-1 -0.34 -1.2
2 4 6 8 10 12 14 1 2 3 4 5 6
t (Å) Number of Fe layers

Fig. 18 Left: calculated total anisotropy energy ΔE of Con Pdm multilayers with (111)-oriented
fcc structure as a function of the thickness t of the magnetic Co layers. Corresponding experimental
data for the product of the anisotropy energy density K and Co thickness t are shown for
polycrystalline films deposited at two different temperatures (triangles up and down). (All data
taken from [163]) Right: SOC-induced (ΔESOC ; circles) and dipole-dipole (ΔEdip ; triangles)
contributions to the total anisotropy energy (ΔE; squares) for the magnetic surface layer system
Fen /Au(001) as a function of the number n of Fe layers. (All data taken from [155])

cases that the classical treatment on the basis of the dipole-dipole interaction leads
to results for ΔEdip (m̂, m̂ ) that are very close to those of a coherent quantum-
mechanical calculation that accounts also for the Breit interaction.
Starting from the 1950s, compounds of rare-earth (RE) with 3d transition metal
(TM) elements, as, for example SmCo5 , or Nd2 Fe14 B, attracted much attention
because of their strong magnetic anisotropy. In these materials, the MCA is
primarily determined by the RE sublattice, while the TM sublattice is responsible
for the magnetic ordering [167]. For that reason, the simplified two-sublattice
Hamiltonian

H = Hd + HCEF
f
+ Hex
d-f
(71)

is often used to discuss their properties, where Hd and HCEF f characterize the TM
and RE, respectively, sublattices while Hex describes the exchange interactions
d-f

between the two. Within the single-ion model, HCEF f accounts for the interaction
of the aspherical 4f-charge with the crystalline electric field (CEF). Due to strong
spin-orbit interaction for the 4f-electrons, rotation of the magnetization leads to a
rotation of their aspherical charge cloud. This in turn results in a dependency of the
electrostatic energy on the orientation of the 4f-magnetic moment as described by
the Hamiltonian [168, 169, 170]


HCEF
f
= n θJ n r
4f On
4f .
Am n m
(72)
n,m
4 Electronic Structure: Metals and Insulators 229

Here Am n are crystal field parameters for the angular momentum quantum numbers
n and m determined by the charge contribution in the system excluding the 4f-
electrons, θJ n are Stevens’ factors depending on the total angular momentum
quantum number J , Onm
4f are the expectation values of the Stevens’ operators, and
r n
4f are the expectation values of r n calculated for the 4f-states of the RE atom.
As the quantities θJ n and Onm
4f are all tabulated, calculation of the crystal field
n together with r
4f allows to fix HCEF and with this to determine
parameters Am n f

the corresponding phenomenological anisotropy constants Ki [168, 169].


While the first calculations in the field have been done adopting a spherical
approximation for the potential [171,172], later work clearly demonstrated the need
to use a nonspherical potential. Such calculations have been performed, for example,
by Richter et al. [173] on SmCo5 representing the itinerant s-, p-, and d-electrons via
band states, while the localized Sm 4f-states are treated within the atomic like so-
called open shell scheme. Hummler and Fähnle report on corresponding calculations
on the CEF parameters for the whole RECo5 series with RE=Ce . . . Yb [170].
Their results for A02 r 2
4f and A04 r 4
4f are plotted in Fig. 19 in comparison with
experiment. This type of calculations on bulk materials led in general to satisfying
agreement with experiment and clearly showed that the naive point charge model
is completely inadequate for an estimate of the CEF parameters: point charges
chosen according to the chemical valency of the elements are much too high when
compared to ionic charges obtained from self-consistent calculations. In addition,
it turned out that the parameters Am n are determined by about 80% by the charge
distribution on the RE site while the point charge model assumes a lattice of ionic
point charges surrounding the RE site.
Corresponding work has also been performed in order to investigate the magnetic
anisotropy at the surface or interface of RE-based compounds. Calculations of

0 0

-100
-10
A2 < r >4f (K)

A4 < r >4f (K)

-200
-20
2

-300
-30
0

-400

-500 -40

Ce Nd Sm Gd Dy Er Yb Ce Nd Sm Gd Dy Er Yb
Pr Pm Eu Tb Ho Tm Pr Pm Eu Tb Ho Tm

Fig. 19 Comparison of theoretical and experimental values for A02 r 2


4f (left) and A04 r 4
4f
(right) parameters for the series of RECo5 compounds with RE = Ce . . . Yb. The full circles (full
squares) are theoretical results for the experimental lattice parameters (for the lattice parameters
fixed to those of GdCo5 ). Experimental values are shown as open squares and crosses. (All data
taken from [174])
230 H. Ebert et al.

A02 r 2
have been done, for example, for the Nd sites of the (001)-surface of
Nd2 Fe14 B [175]. It turned out that the sign of A02 r 2
depends on the positions of
the Nd atoms in the unit cell supplying this way an explanation for the different
coercivity of crystalline and sintered Nd2 Fe14 B. Similar calculations have been
performed also to investigate the impact of Dy impurities on the coercivity of
Nd2 Fe14 B [176]. From these, it was found that the parameter A02 r 2
for Dy
atoms in the surface region of Nd2 Fe14 B also may have a positive or negative sign
depending on its position, leading finally to a decrease of the coercivity of sintered
samples.
P. Novák et al. introduced a scheme for the calculation of the crystal field
parameters that avoids the assumption of an inert 4f-charge cloud and allows for
the hybridization of the 4f-states with the surrounding electronic states [177].
The approach is based on a local Hamiltonian represented in the basis of Wannier
functions and expanded in a series of spherical tensor operators. Applications to RE
impurities in yttrium aluminate showed that the calculated crystal field decreases
continuously as the number of 4f-electrons increases and that the hybridization of
4f-states with the states of the oxygen ligands is important. This method has been
successfully applied also to calculate crystal field parameters for RE impurities in
LaF3 [178].
Dealing with ferrimagnetic materials composed of several, inequivalent mag-
netic sublattices, calculating the magnetic anisotropy may become more compli-
cated as an additional canting between the sublattices introduced by an external
field may play a significant role and should be taken into account to get a
reasonable agreement with experiment. This was demonstrated for the RE-TM
ferrimagnet GdCo5 [179], where the authors report a first-principles magnetization-
versus-field (FPMVB) approach giving temperature-dependent magnetization as
a function of an externally applied magnetic field in excellent agreement with
experiment.

Excitations

Many dynamical as well as finite-temperature properties of the magnetization of a


solid can be understood and described on the basis of magnetic excitations. In the
low-energy, small-wave vector regime, one has to deal with the collective magnon
excitations that can be investigated by various techniques. An approximate approach
builds on the use of calculated exchange coupling parameters in combination with
the so-called rigid spin approximation. More accurate results can be expected from
self-consistent spin-spiral or frozen-magnon calculations that also allow exploring
the magnetic phase space in an efficient way. Both approaches, however, do not give
access to single-particle or Stoner excitations. On the other hand, using the concept
of the dynamical susceptibility depending on frequency and wave vector, a coherent
description of magnon and Stoner excitations is achieved.
4 Electronic Structure: Metals and Insulators 231

Magnon Dispersion Relations Based on the Rigid Spin


Approximation

When considering the magnetization dynamics of solids, one usually assumes the
magnetization to be collinear inside an atomic cell i oriented along the common
direction m̂i implying a coherent rotation of the magnetization within the cell during
progress of time (rigid spin approximation (RSA)) [144]. As a consequence, the
equation of motion for the magnetization can be replaced by the equation of motion
for the local magnetic moments mi = mi m̂i that can be written as [144, 180, 181]:

d 2μB 1 ∂E
m̂i = − × m̂i , (73)
dt h̄ mi ∂ m̂i

where the right-hand side represents the torque acting on the magnetic moment mi .
Making use of the harmonic approximation for the energy, Eq. (73) yields for the
spin waves uλν (q) = uλν eiqR n with wave vector q the following eigenvalue problem
[181]:

2μB  νν 
h̄ ωλ (q) uλν = J (q) uλν , (74)
mν 
ν

for solids with translational symmetry. Here the eigenvectors uλν numbered by the
index λ represent small deviations of magnetic moments from the direction of the

ground state and the J νν (q) are the Fourier transforms of the interatomic exchange
coupling parameters with ν labeling the basis atoms within a unit cell.
Solution of the eigenvalue problem in Eq. (74) obviously yields the frequencies
ωλ (q) of the various collective spin-wave eigenmodes that can be compared with
magnon excitation energies as deduced, for example, from neutron scattering.
Corresponding results obtained for Fe and Ni are given in Fig. 20 in comparison
with experiment. Although good agreement between theory and experiment is
achieved, one has to stress that the theoretical results depend on the method used to

calculate the J νν (q) parameters. The data shown by a solid line were obtained using
the exchange coupling parameters Jij calculated on the basis of the Lichtenstein
formula Eq. (52) [182]. In the case of a lattice with one atom per unit cell, the
magnon energy spectra E(q) possess only a single branch. Therefore, the bcc-
Fe and fcc-Ni magnon spectra in Fig. 20 could be obtained by a simple Fourier
transformation [182]

4μB 
E(q) = J0j (1 − eiq·R j ). (75)
m
j

The minima of E(q) for bcc-Fe along the Γ − H and H − N directions to be


seen in Fig. 20 are so-called Kohn anomalies which occur due to long-range RKKY
interactions. It turned out that these minima appear only if the summation in Eq. (75)
232 H. Ebert et al.

600 Expt. 1 500


Fe Expt. 2 Ni
500 Halilov et al.
Pajda et al. 400

E(q) (meV)
E(q) (meV)

400
300
300
200
200 Expt
Halilov et al.
100 Pajda et al.
100

0 0
Γ N P Γ H N L Γ X W K Γ

Fig. 20 Magnon dispersion relations for bcc-Fe (left) and fcc-Ni (right) along high-symmetry
directions in the Brillouin zone, in comparison with experiment (open symbols [182]). Solid lines
represent the results by Pajda et al. [182], while full circles show the results by Halilov et al. [180]

700
Co 30 Gd Expt
600 Theory
25
500
E(q) (meV)

E(q) (meV)

400 20

300 15

200 10

100 5

0 0
Γ M K Γ A L Γ M K Γ A

Fig. 21 Left: magnon dispersion relation for hcp-Co along high-symmetry lines in the Brillouin
zone [180]. Right: magnon dispersion relation for hcp-Gd (full lines) in comparison with
experimental data. (All data taken from [183])

is performed over a sufficiently large number of atom shells around the central
atomic site with index 0.
As an example for a lattice with a multiatom basis, Fig. 21 (left) displays the
magnon spectrum calculated for hcp-Co [180]. As there are two atomic sites in the
unit cell, solving the eigenvalue problem Eq. (74) leads to two magnon branches.
A similar approach was applied to hcp-Gd [183]. The corresponding experimental
data shown in Fig. 21 (right) were obtained at T = 78 K. This was accounted for
in the calculations within the RPA (see section “Methods Relying on the Rigid Spin
Approximation”) leading to a simple rescaling of the magnon energies proportional
to the temperature-dependent average magnetization.
4 Electronic Structure: Metals and Insulators 233

Spin Spiral Calculations

Usually calculations of the electronic structure for magnetic systems are performed
assuming a collinear spin-magnetic structure and using the smallest unit cell
corresponding to the space group of the system. However, this configuration does
not have to correspond to the ground state of the system. A possible way to
search for the proper magnetic ground state is to consider incommensurate spin-
spiral configurations. Furthermore, within the adiabatic approximation, spin spirals
can be seen as a representation of transverse spin fluctuations. Therefore, self-
consistent calculations on static spin spirals or so-called frozen magnons give
access to the energies of spin-wave excitations that can be used in particular to
investigate the finite-temperature magnetism. Considering a corresponding spin
spiral characterized by the wave vector q and the tilt angles θν and φν , the variation
of the spin-magnetic moment mnν from site to site may be expressed via:

mnν = mν [cos(q · R n + φν ) sin θν , sin(q · R n + φν ) sin θν , cos θν ], (76)

where ν labels the atomic site in the unit cell located at lattice vector R n .
Because of broken translational and rotational symmetry, the presence of a spin
spiral in principle implies an increased unit cell compared to a collinear spin
configuration. However, as shown by Brinkman and Elliot [184, 185] as well as
Herring [186], one can make use of the fact that a spin-spiral structure characterized
by the wave vector q is invariant with respect to a so-called generalized translation:

Tn = {α(φ)|αR |t n }, (77)

if spin-orbit coupling is neglected. Here, the vector t n specifies a spatial translation


combined with a spatial rotation αR and a spin rotation about the ẑ axis by the angle
α(φ) = α(q · t n ). This property allows to formulate the generalized Bloch theorem
[187]:

Tn ψj k (r) = e−ik·t n ψj k (r), (78)

that specifies the behavior under a generalized translation for the two-component
eigenfunctions ψj k of a Hamiltonian with a noncollinear spin-dependent potential
of the form (see also Eq. (11)):
 
 q†

Vnν (r) 0 q
V (r) = Unν (θν , φν ) ↓ Unν (θν , φν ). (79)

0 Vnν (r)

Here n specifies the Bravais lattice vector R n , ν gives the position ρ ν of an atom
q
in the unit cell, and Unν is a spin-transformation matrix that connects the global
frame of reference of the crystal to the local frame of the atom site at R n + ρ ν
that has its magnetic moment mnν tilted away from the global z-direction. The
234 H. Ebert et al.

x ϕ=qR
m
θ
z
y

Fig. 22 Geometry of a spin spiral with the wave vector q along the z-direction

q
transformation Unν is characterized by the Euler angles θnν and φnν as it is shown
in Fig. 22. Assuming a collinear alignment of the spin density within the atomic cell
at (n, ν), it is natural to use a local frame of reference with its z-axis oriented along
q
mnν . The corresponding transformation matrices Unν occurring in Eq. (79) can be
q
written as a product of two independent rotation matrices Unν = Un (θν , φν , q) =
Uν (θν , φν ) UqR n , where the matrix UqR n depends only on the translation vector R n
[187]:

  i
 i

q cos θ2ν sin θ2ν e 2 φν 0 e 2 q·Rn 0
Unν = . (80)
− sin θ2ν cos θ2ν 0e − 2i φν
0e − 2i q·Rn

Results for the wave vector-dependent energy and spin-magnetic moments per
atom obtained by Uhl et al. [188] from corresponding spin-spiral calculations
for γ -Fe are shown in Fig. 23. This work was based on the LSDA and used the
augmented spherical wave (ASW) band structure method in combination with the
atomic sphere approximation (ASA). In addition, noncollinearity within an atomic
cell was neglected. The investigations on γ -Fe by Kurz et al. [189], on the other
hand, avoided these simplifications by the use of the LAPW band structure method.
The noncollinear magnetic structure was imposed by a constraining magnetic field
applied to the magnetic moments of the atoms. Furthermore, the GGA was used
for the exchange-correlation potential. In spite of the various technical differences
between the two studies, the results shown in Fig. 23 agree fairly well and justify
the approach used by Uhl et al. as well as many others.
The implementation of the spin-spiral method within the KKR band structure
method allows dealing not only with ordered materials but also with random alloys
[190]. Figure 24 gives as an example the energy (left) and individual spin moments
(right) for a spin-spiral magnetic structure in Fe0.5 Mn0.5 as a function of the wave
vector q directed along the [001] direction. One can see a transition from the
antiparallel alignment of the Fe and Mn magnetic moments at small q to a parallel
alignment when q approaches the boundary of the Brillouin zone at |q| = 2π/a.
As it is seen from the left part of Fig. 24, the latter magnetic configuration is
energetically more stable.
Apart from exploring the magnetic phase space by performing self-consistent
spin-spiral calculations, the technique can also be used to get access to magnon
4 Electronic Structure: Metals and Insulators 235

0 2

-10 Kurz et al.


Uhl et al. 1.5
-20
E(q) (meV)

mspin (μB)
-30 1

-40 Kurz et al.


0.5 Uhl et al.
-50

-60 0
Γ X W Γ X W

Fig. 23 The energies of the spin-spiral structure with respect to the energy of the FM state (left)
and spin-magnetic moments per atom (right) calculated for γ -Fe for the wave vector q varying
along Γ − X − W in the Brillouin zone. Open diamonds represent the results obtained by Uhl et al.
[188] while full circles represent the results obtained by Kurz et al. [189]. (All data taken from
[188] and [189])

3
0.2 MFe
Mn and Fe parallel 2.5 MMn
2 Mtotal
Espin spiral (eV)

1.5
Mspin (μB)

0.1
1
0.5
Mn and Fe
antiparallel 0
0
-0.5
-1
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
qz qz

Fig. 24 Left: energy of a spin spiral in a Fe0.5 Mn0.5 alloy calculated for the wave vectors q =

a (0, 0, qz ) along the [001] direction. Right: local magnetic moments on Fe and Mn atoms as a
function of the wave vector q. (All data taken from [190])

excitation energies h̄ω(q). In the case of simple lattices, the energy ΔE(q, θ ) of a
spin spiral with wave vector q and tilt angle θ can be used directly to get h̄ω(q)
from the expression [191]:

4 ΔE(q, θ )
h̄ω(q) = lim . (81)
θ→0 m sin2 θ

This approach was used, for example, by Halilov et al. [180] to calculate the magnon
energy spectra for Fe and Ni represented in Fig. 20. Obviously, the results are in a
reasonably good agreement with those of Pajda et al. [182] that are based on a
236 H. Ebert et al.

calculation of the real space exchange coupling parameters Jij via the Lichtenstein
formula Eq. (52). As one notes, the minima for Fe along the Γ − H and H − N
directions are given by both approaches. However, the magnons obtained by the
spin-spiral calculations are softer, because of the self-consistent relaxation within
the electronic structure calculations.
As another example, Fig. 25 represents spin-wave dispersion curves obtained for
the full-Heusler alloys Cu2 MnAl, Pd2 MnSn, Ni2 MnSn, and the L12 -type ferromag-
net MnPt3 [192]. The simplified approach used in this work did not fully account for
the magnetic sublattices of the investigated systems providing for that reason only
the first magnon branch. Nevertheless, this already led to values for the Curie tem-
perature calculated within the RPA approach (see section “Methods Relying on the
Rigid Spin Approximation”) in reasonable agreement with experiment. As a general
trend, one can see in Fig. 25 that the calculated magnon energies h̄ω(q) are too high
when compared with experiment. The authors attribute this to the treatment of elec-
tronic correlations on the basis of the GGA. In fact, previous work on the series of
Heusler alloys Co2 Mn1−x Fex Si [62] clearly demonstrated the impact of correlation
effects by comparison of results based on the LSDA, LSDA+U, and LSDA+DMFT.

200 Expt (4.2 K)


Cu2MnAl Pd2MnSn Expt (50 K)
Theory 60 Theory
150
E(q) (meV)
E(q) (meV)

40
100

20
50

[100] [110] [111] [100] [110] [111]


0 0
Γ X Γ L Γ X Γ L
80 200
Ni2MnSn Theory MnPt3 Expt (80 K)
Expt (50 K) Theory
60 150
E(q) (meV)

E(q) (meV)

40 100

20 50

[100] [110] [111] [100] [110] [111]


0 0
Γ X Γ L Γ X M Γ R

Fig. 25 Calculated (solid lines) spin-wave dispersion curves h̄ω(q) in the first Brillouin zone
along high-symmetry directions for the L21 -type full-Heusler and L12 -type ferromagnets. As indi-
cated, the experimental data stem from neutron diffraction measurements at various temperatures.
(All data taken from [192])
4 Electronic Structure: Metals and Insulators 237

The use of the spin-spiral technique for the calculation of the full magnon energy
spectrum in case of complex compounds was demonstrated by Şaşıoğlu et al. [193].
In this case, a set of spin-spiral calculations is required to obtain all exchange

coupling parameters J νν (q) that enter the eigenvalue problem Eq. (74).

Excitation Spectra Based on the Dynamical Susceptibility

Despite the many successful applications of the adiabatic approach for the investi-
gation of spin-wave excitations, one has to stress that it has severe limitations. In
particular it can be applied only to systems for which single-particle or the so-called
Stoner excitations can be neglected [194, 195]. Figure 26 gives a simplified picture
of the exchange-split band structure of an itinerant ferromagnet in the vicinity of its
Fermi level. Excitation of an electron from an occupied majority-spin state below
the Fermi level to an empty minority state may not only be associated with a spin
flip but also with a change of the electronic wave vector. As it is visualized in the
right panel of Fig. 26, these Stoner excitations lead to a broad continuum that in
general overlaps and hybridizes with the discrete magnon dispersion spectrum. For
this and other reasons, a more sophisticated description of spin-wave excitations was
worked out making use of the linear response formalism within the framework of
time-dependent density functional theory [196,197,87,198]. This allows expressing
the magnetization Δmi (r, q, ω) induced by a magnetic field B(r, q, ω), with its
time and spatial dependency expressed by the wave vector q and frequency ω,
respectively, in terms of a corresponding susceptibility tensor [2]:


Δmi (r, q, ω) = d 3 r  χ ij (r, r  , q, ω) B j (r  , q, ω). (82)
j Ω

Δk = q Minority-spin
Excitation energy, ε
Electron energy, E

EF
ΔE = ε δ Stoner continuum
U

U
Magnons
Majority-spin
0 δ
Electron wave vector, k Excitation wave vector, q

Fig. 26 Left: schematic representation of Stoner excitations in an itinerant ferromagnet. A


majority electron is excited from an occupied state below the Fermi level to an unoccupied minority
state above the Fermi level. Right: continuum of Stoner excitations for a metallic ferromagnet
238 H. Ebert et al.

Within spin-density functional theory and making use of circular coordinates, one
may write in particular for the transverse susceptibility tensor element χ ± a Dyson-
like equation [87]:

χ ± (r, r  , q, ω) = χ0± (r, r  , q, ω)



+ d 3 r  χ0± (r, r  , q, ω) Kxc (r  ) χ ± (r  , r  , q, ω), (83)
Ω

where χ0± is the unenhanced susceptibility, Kxc (r) = Bm(r) xc (r)


is the exchange-
correlation kernel function, and Bxc (r) and m(r) are the local exchange-correlation
field and magnetization, respectively. As discussed, for example, by Bruno [199] as
well as by Katsnelson and Lichtenstein [200], it is the second term in Eq. (83) that
gives rise to the enhancement of the transverse susceptibility.
The dynamical susceptibility gives not only access to the energetics of magnetic
excitations but also to their lifetime characterizing this way the dissipation of the
energy. Outside the Stoner continuum, the loss tensor associated with χ + shows
peaks at frequencies corresponding to the excitation of spin waves. Inside the Stoner
continuum, these show a finite width due to the hybridization with the Stoner
excitations. In this case, one has approximately:

Aλ (q)
χλ+ (q, ω) ≈ , (84)
(ω − ω0λ (q))2 + βλ (q)2

with the amplitude Aλ (q), the spin-wave energy ω0λ (q), and inverse lifetime βλ (q).
As an example, Fig. 27 shows the magnon dispersion curves ω0 (q) together with
the corresponding broadening for the magnon states as deduced from the dynamical
susceptibility as calculated for bcc-Fe and fcc-Co [203]. As one can clearly see
from the given width, Stoner excitations have only a very small influence on the
long-period magnons.
Results for ω0λ (q) and βλ (q) for the Heusler alloy Co2 MnSi that has three
magnetic sublattices are shown in Fig. 28. The corresponding eigenvectors (EV)

Fig. 27 Spin waves of bcc-Fe (left) and fcc-Co (right). Solid circles correspond to ω0 (q), while
the error bars denote full width at half maximum of the peak. Solid line denote spin-wave energies
obtained using the magnetic force theorem [203]
4 Electronic Structure: Metals and Insulators 239

800 150 EV 2
EV 3

600
ω(q) (meV)

β(q) (meV)
EV 1 100
400 EV 2
EV 3

200 50

0
Γ [ξ00] X K [ξξ0] Γ [ξξξ] L Γ [ξ00] X K [ξξ0] Γ [ξξξ] L

Fig. 28 Energies ω0λ (left) and inverse lifetimes βλ (middle) of three spin-wave modes in
Co2 MnSi together with the corresponding eigenvectors (right); arrows indicate the orientations
of the magnetic moments. The basis atoms are Co at (1/4, 1/4, 1/4)a and (3/4, 3/4, 3/4)a and Mn
at (1/2, 1/2, 1/2)a. The parameter β1 of EV 1 does not exceed 5 meV and is not shown. (All data
taken from [87])

of the resulting spin-wave modes are given on the right-hand side of the fig-
ure. The acoustic mode that is lowest in energy has a vanishing energy for
q = 0, and its value for βλ is very small (therefore not shown in Fig. 28).
The optical modes, on the other hand, appear at higher energies where the
continuum density is appreciable. Accordingly, their inverse lifetime βλ is quite
large and depends strongly and non-monotonously on the wave vector q. For
the Heusler alloy Cu2 MnAl, only one magnetic sublattice has to be consid-
ered, and accordingly, there is only one acoustic spin-wave mode. In contrast
to Co2 MnSi, a more pronounced damping is found in this case. The influence
of the Stoner excitations can also be seen for the spin-wave energies. Calcu-
lating these by use of the so-called adiabatic approach [202] that neglects the
hybridization, the spin-wave energies are higher and in less good agreement with
experiment [202].
Comparable studies based on the dynamical susceptibility were done (i) to
investigate the Landau damping in Fe(100) and Fe(110) films and the effect of
the substrate on this [201], (ii) to study the Landau damping of spin waves
and large Rh moments induced by the AFM magnons in FeRh [203], and
(iii) on acoustic magnons in the long-wavelength limit in order to analyze the
Goldstone violation in many-body perturbation theory [204]. The concept of
the dynamic spin susceptibility has been applied also to paramagnetic systems
at finite temperatures by Staunton et al. [205, 206]. Due to the use of the
multiple scattering formalism, investigations on alloys could be made, for exam-
ple, on paramagnetic Cr0.95 V0.05 and antiferromagnetic Cr0.95 Re0.05 above the
Néel temperature TN . While the work sketched here was primarily based on
the linear response formalism applied within the framework of time-dependent
density functional theory, similar work on quasiparticle and collective electronic
excitations in solids was done using techniques from many-body perturbation
theory [207].
240 H. Ebert et al.

Finite-Temperature Magnetism

Dealing with the impact of finite temperatures in a quantitative way is a big


challenge for theory. Accordingly, many different techniques on various levels of
sophistication are in use for that purpose. Most of these employ the adiabatic
approximation that decouples the electronic and magnetic degrees of freedom.
One type of such approaches, that proved to be astonishingly successful for
many situations, starts from the properties of low-energy magnetic excitations by
calculating real-space exchange coupling parameters or the energies of spin spirals.
In a second step, this information is used in combination with classical statistical
methods including in particular the Monte Carlo method to deduce temperature-
dependent magnetic properties. More advanced schemes, however, are based on a
coherent description of the electronic structure and statistics. While the disordered
local moment (DLM) method still relies on the adiabatic approximation, this does
not apply, for example, to the functional-integral method or various many-body
approaches used within the dynamical mean field theory (DMFT) that account for
finite temperature in a coherent way.

Methods Relying on the Rigid Spin Approximation

Within standard Stoner theory, a spin-dependent but collinear electronic structure is


assumed, and finite temperatures are accounted for only via the Fermi distribution
function. Accordingly, the resulting critical temperatures are much too high. More
successful approaches to deal with magnetism at finite temperatures, on the other
hand, allow for transverse spin excitations. A simple model accounting for this was
suggested already by P. Weiss who considered a magnet as a system of localized
magnetic moments that order spontaneously due to an effective molecular or Weiss
field hW (T ) = w m(T ) n̂ that depends on the average magnetic moment m(T ) on
an atomic site. The factor w is the molecular or Weiss field constant:

3kB TC
w= , (85)
m20

which is determined by the magnetic moment m0 = m(0) at T = 0 K and the critical


temperature TC . Quite general, the temperature-dependent magnetic moment m(T )
along n̂ is determined by the statistical average over all possible orientations ê with
the probability distribution for the local magnetic moments given by:

e−β hW n̂·ê
P n̂ (ê) =  , (86)
d ê e−β hW n̂·ê

where hW = w m(T ) and β = 1/(kB T ). The various techniques discussed in


section “Exchange Coupling Parameters” allow to deduce the Weiss or mean field
4 Electronic Structure: Metals and Insulators 241

constant from the calculated exchange coupling parameters Jij . Within this mean
field approach (MFA), application of classical spin statistics leads to:

2  2
TCMFA = J0j = J0 , (87)
3kB 3kB
j =0

for MFA in case of an elemental ferromagnet, with J =


 the Curie temperature TC 0
j =0 J0j [126].
A more accurate approach to deal with finite-temperature magnetism is provided
by the random-phase approximation Green function (RPA-GF) method which also
can be based on a combination of the Heisenberg model and SDFT calculations
(see section “Exchange Coupling Parameters”). The decoupling scheme suggested
by Tyablikov leads to an approximate expression for the one-particle Green magnon
function [208, 209]:

1  1
Gm (z) = , (88)
N q z − E(q)

with E(q) the magnon energies that allows expressing the critical, i.e., Curie or
Néel, respectively, temperature TcRPA as [182]:

1 6
RPA
= − lim Gm (z), (89)
k B Tc z→0 m

in case of a ferro- or antiferromagnet, respectively. The MFA approach accounts


for all spin-wave excitations with the same weight leading in general to an
overestimation of the Curie temperature. The RPA-GF approach, on the other hand,
accounts in particular for the low-energy excitations in a much more adequate
manner. As a consequence, more accurate results for the critical temperature are
normally obtained this way when compared with experiment. This behavior has
been found within numerical work on the elemental ferromagnets Fe, Co, and Ni
[182] but also for other systems, as, for example, the Heusler alloys (Ni,Cu)2 MnSn,
(Ni,Pd)2 MnSn [210], NiMnSb, CoMnSb, Co2 MnSi, and Co2 CrAl [193], for Gd-
based intermetallic compounds GdX (X = Mg, Rh, Ni, Pd) [211] as well as for
zincblende half-metallic ferromagnets GaX X = N, P, As, Sb [212]. The exchange
coupling parameters in these works have been obtained either from spin-spiral
calculations [193,212] or in real space, on the basis of the force theorem [182,211].
In all cases, reasonable agreement with experiment could be achieved, except for
fcc-Ni, a system for which the application of a Heisenberg Hamiltonian with fixed
spin moments seems questionable.
Corresponding investigations on the spin-wave spectra and Curie temperatures
have also been performed for L21 -type full-Heusler FM alloys and L12 -type XPt3
alloys [192]. It was found that the Curie temperatures are in good agreement with
experiment when the Stoner gap was large enough so that the magnon regime is
well separated from the Stoner excitations. In this case, single-particle spin-flip
242 H. Ebert et al.

Stoner excitations make only a small contribution to the excitation spectra at


low energies, so that magnon excitations make the dominating contribution to
thermodynamics. It is interesting to note that the RPA-GF formalism fulfills the
Mermin-Wagner theorem [213], i.e., in the absence of magnetic anisotropy it leads
for two-dimensional systems to TC = 0 K [214]. Corresponding studies by Pajda
et al. on the Curie temperature in Fe and Co films that included a finite magnetic
anisotropy led to an oscillating behavior as a function of film thickness [127]. These
oscillations could be ascribed to the oscillating behavior of the exchange coupling
parameters due to thickness-dependent quantum well states. An extension of the
RPA-GF method (RPA-CPA) to deal within the CPA also with disordered alloys
was worked out by Bouzerar and Bruno [215].
Starting from the Stratonovich-Hubbard functional integral method [83], Kübler
[191] derived an expression for the critical, i.e., Curie or Néel, respectively,
temperature:
⎡ ⎤−1
2 2⎣1  1 ⎦
k B Tc = L (90)
3 ν ν N q,n jn (q)

that is applicable to systems with several atoms per unit cell. Here Lν is the so-
called local moment of atom ν that should be found in a self-consistent way [191].
However, in the case of well-localized magnetic moments, values of mν at T = 0
K can be used as a reasonable approximation [216]. Finally, jn (q) in Eq. (90)
is the exchange function after diagonalization of jνν  (q) [191]. This expression
accounts in particular for the fact that different atomic types have in general different
moments, while the expressions in Eqs. (88) and (89) that are often applied to
multicomponent systems are based on the assumption that Lν = m0 with m0 , the
average saturation moment. As is demonstrated by the results in Table 2, Eq. (90)
gives in general results for the critical temperature that are in good agreement with
experiment.
An important alternative to the RPA-GF scheme for dealing with magnetic
properties at finite temperatures is provided by the Monte Carlo method that is
used to deal with the statistical aspect of the problem. Corresponding work again
is in general based on the Heisenberg Hamiltonian (see Eq. (51)) with the exchange
parameters calculated in an ab initio way. Numerous successful applications have
been done, both for ordered compounds [218] and for disordered alloys such as
diluted magnetic semiconductors like Ga1−x Mnx As [135, 219, 220] or Heusler
alloys [129, 221]. In most cases, good agreement with RPA-based results as well as
with corresponding experimental data for the critical temperature could be achieved.

Methods Accounting for Longitudinal Spin Fluctuations

Despite the good results often obtained via the RPA and MC methods, one has
to stress that they are based on a classical spin Hamiltonian and therefore are
4 Electronic Structure: Metals and Insulators 243

Table 2 Structure, magnetic moments on M1 and M2 sublattices, as well as critical Curie or Néel
temperature calculated via the MC (TcMC ), RPA (TcRPA ), or MFA (TcMFA ) approaches in comparison
with experiment. (All data taken from [191] and [217])
Exp
System Structure M1 (μB ) M2 (μB ) TcMC/RPA (K) TcMFA (K) Tc (K)
Fe bcc 2.330 – 1060MC 1460 1043
Co fcc 1.410 – 1080MC 1770 1388
Ni fcc 0.630 – 510MC 660 633
FeNi CuAu 2.551 0.600 972RPA 1130 790
CoNi CuAu 1.643 0.673 1149RPA 1538 1140
FeNi3 AuCu3 2.822 0.588 986RPA 1290 870
CoNi3 AuCu3 1.640 0.629 733RPA 925 920
NiMnSb C1b 3.697 0.303 968RPA 1281 730
Mn2 VAl L21 −0.769 1.374 580RPA 663 760
Co2 FeSi L21 2.698 1.149 1058RPA 1267 1100
Mn3 Al L21 −2.258 1.128 196RPA 342 –
Mn3 Ga L21 −2.744 1.363 314RPA 482 –
Mn3 Ga DO22 −2.829 2.273 762RPA 1176 –
RhMn3 AuCu3 3.066 – 1059RPA – 855

suitable only for systems with their magnetic moments depending only weakly on
the temperature. As the latter assumption is not always fulfilled, an extension of
the Heisenberg Hamiltonian was suggested that is meant to account for longitudinal
fluctuations [222, 217] that express the total energy by an expansion in even powers
of the magnetic moments per atom m:
 
E(M, q, θ ) = An m2n + Jn (q, θ ) m2n . (91)
n n

Here the functions Jn (q, θ ) are proportional to the energy difference between the
ferromagnetic and the spin-spiral states specified by wave vector q and tilt angle θ .
Monte Carlo simulations for bcc-Fe, fcc-Co and fcc-Ni performed on this basis
[217] led to a rather good agreement with experiment as one can see from Table 2.
Another model to account for temperature-induced longitudinal spin fluctuations
was suggested by Ruban et al. [223] that also led for Fe and Ni to a rather realistic
description of the magnetic properties at finite temperatures.
An important class of materials, for which longitudinal spin fluctuations are
of great importance, are alloys and compounds composed of magnetic and other-
wise nonmagnetic elements. Such systems exhibit so-called covalent magnetism
[224, 225], i.e., the magnetization of the nonmagnetic component is caused by
the spontaneously magnetized atoms via a spin-dependent hybridization of the
electronic states. Ležaić et al. [129, 221], for example, emphasized the need
to account for longitudinal fluctuations of the magnetic moment induced on the
Ni atoms for a proper description of the temperature dependence of the spin
polarization at the Fermi energy EF when performing Monte Carlo investigations
244 H. Ebert et al.

on the half-metallic Heusler alloy NiMnSb. For that purpose, they used an extended
Heisenberg Hamiltonian:

1 
Hext = Jij mi · mi + (a m2i + b m4i ), (92)
2
ij i∈N i

that allows accounting for transverse and longitudinal magnetic fluctuations con-
nected with the temperature-dependent induced magnetic moment on the Ni atom.
The Heusler alloy NiMnSb was also investigated by Sandratskii et al. [212] using
the spin-spiral approach and treating the magnetic moment of Ni atom as being
induced.
Mryasov et al. [226] found that the induced magnetic moment of Pt plays a
crucial role for the magnetic anisotropy of FePt at finite temperatures. To account
for this, a renomalization of the Fe-Fe exchange interactions according to:

 2
χPt 
J˜ij = Jij + I Pt
Jiν Jνj (93)
m0Pt ν∈Pt

was suggested. Here I Pt characterizes the local exchange interaction of the Pt


atoms, m0Pt is the Pt magnetic moment in the ordered ferromagnetic state, and
χPt is the partial spin susceptibility of Pt. The impact of a renormalization of
the Fe-Fe interactions due to the induced Rh magnetic moment has also been
demonstrated for the stabilization of the ferromagnetic state and for the control of
the antiferromagnet-ferromagnet phase transition of FeRh [227].
Another scheme to account within Monte Carlo simulations for the impact of
the induced magnetic moments on nonmagnetic alloy components leading to a
renomalization of the exchange interactions between the magnetic components
was worked out by Polesya et al. [128]. Figure 29 (left) shows corresponding
results for the Curie temperatures of the ferromagnetic alloy Fex Pd1−x that are in
very good agreement even for the Pd-rich side of the alloy system. Corresponding
calculations have been done for FeRh to investigate its antiferromagnet-ferromagnet
phase transition [229]. The right panel of Fig. 29 shows results of Kudrnovský
et al. [228] for fcc-Ni1−x Pdx . Investigating the finite-temperature magnetism of
various Ni-based transition metal alloys, these authors concluded that Bruno’s
formulation [199] for the renormalized RPA gives the most satisfying agreement
with experiment.

Coherent Treatment of Electronic Structure and Spin Statistics

The methods sketched above that are based on the Heisenberg model obviously
have problems to account for longitudinal spin fluctuations. In addition, they consist
in an incoherent combination of electronic structure calculations and classical
4 Electronic Structure: Metals and Insulators 245

500 Theory RPA


Expt. 1 600 MFA
Expt. 2 rRPA
400 Expt. 3 Expt.
Expt. 4
400

T (K)
T (K)

300

200
200
100

0 0
0 0.05 0.1 0.15 0.2 0 0.2 0.4 0.6 0.8 1
xFe xNi

Fig. 29 Calculated Curie temperature TC of disordered alloys for fcc-Fex Pd1−x (left) using the
Monte Carlo method and fcc-Nix Pd1−x (right) using the MFA, RPA, and renormalized RPA (rRPA)
approaches, in comparison with experiment. (All data taken from [128] and [228])

statistics. These problems can be avoided by the use of the disordered local moment
(DLM) method [230] that deals with the temperature-dependent magnetization
within self-consistent electronic structure calculations. The slow dynamics of the
magnetic moments spatially localized on the atoms when compared to the fast
electron propagation and relaxation time scale allows to make use of the adiabatic
approximation. Assuming ergodicity for the system of local magnetic moments, the
time average required to calculate the average magnetization at a given temperature
can be replaced by the average over the ensemble of all orientational configurations
characterized by a set of unit vectors {êi }. Within the DLM theory, this is determined
by the corresponding single-site probabilities P n̂ (êi ) obtained from Eq. (86) for
the Weiss field hW = hn̂ . Following Györffy et al. [230], hn̂ is determined by
approximating the free energy F corresponding to the microscopic Hamiltonian
H of the considered system by the free energy F0 based on the trial Hamiltonian
H0n̂ = j hn̂ · êj with hn̂ = hn̂ n̂ and using the Feynman-Peierls inequality [231]:

F ≤ F0 + H − H0
, (94)

where the canonical distribution Eq. (86) is used to calculate the average. Using the
Weiss field hn̂ as a variational parameter to minimize the right-hand side of Eq. (94)
one is led to [162]:

n̂ 3
h = d êi êi Hn̂
êi , (95)

where Hn̂
êi denotes the average of Hn̂ with the restriction that the orientation
of the moment on site i is fixed to êi [230]. Self-consistent electronic structure
calculations for a given temperature result in a temperature-dependent magnetic
moment that automatically accounts for longitudinal fluctuations. The resulting
246 H. Ebert et al.

magnetization M(T ) = M(T )n̂ can be obtained from



M(T ) = mloc (T ) d êi P n̂ (êi ) n̂ · êi , (96)

with the local moment mloc to be determined self-consistently.


Based on a relativistic formulation of the DLM method (RDLM), Staunton et al.
[162] worked out a scheme to investigate the magnetocrystalline anisotropy at finite
temperature. This implies in particular a corresponding extension of Eq. (68) that
allows calculating the magnetic torque together with the Weiss field. Results for
the temperature-dependent magnetization (M(T )) and uniaxial magnetocrystalline
anisotropy (K(T )) that have been obtained by an application of this approach
to L10 -ordered FePt are shown in Fig. 30 [232]. In line with experiment, the
orientation of the easy axis was found to be perpendicular to the Fe- and Pt-layers for
all temperatures. Also the temperature dependence of the anisotropy constant K(T )
is in good agreement with experiment. The single-ion anisotropy model, on the other
hand, fails to give a correct qualitative description for K(T ). Similar RDLM-based
investigations have been performed also for the L10 -ordered FePd [162]. As in
the case of FePt, the easy axis in FePd is perpendicular to the Fe- and Pd-layers,
with the uniaxial magnetocrystalline anisotropy also showing the scaling behavior
K(T ) ∼ [M(T )/M(0)]2 .
Buruzs et al. [233] used the RDLM method to investigate the temperature-
dependent magnetic properties of Co films deposited on Cu(100) (Con /Cu(100))
with the thickness n of the Co film varying from 1 to 6. The resulting Curie
temperatures are given in the Table 3. In agreement with experiment, it was
found that the magnetization is oriented parallel to the surface for almost all
temperatures below the Curie temperature, except for the system with n = 2.
Based on the calculation of the anisotropy constant, a temperature-induced reversal

0.8
ΔESOC (meV)

-0.5
0.6
M(T)

-1
0.4
-1.5
0.2

0 -2
200 400 600 800 0.2 0.4 0.6 0.8
2
Temperature T (K) (M(T))

Fig. 30 RDLM calculations on FePt. Left: the magnetization M(T ) versus T for the magnetiza-
tion along the easy [001] axis (filled squares). The full line shows the mean field approximation
to a classical Heisenberg model for comparison. Right: the magnetic anisotropy energy ΔESOC
as a function of the square of the magnetization M(T ). The filled circles show the RDLM-based
results, the full line give K(T ) ∼ [M(T )/M(0)]2 , and the dashed line is based on the single-ion
model function. (All data taken from [232])
4 Electronic Structure: Metals and Insulators 247

Table 3 Calculated Curie n 1 2 3 4 5 6


temperatures (K) for
TC 1330 933 897 960 945 960
Con /Cu(100) [233]

of the anisotropy direction from in-plane to out-of-plane was predicted. A more


detailed investigation led to the conclusion that the spin reorientation is driven by a
competition of exchange and single-site anisotropies [18].
Zhuravlev et al. [234] used the RDLM method to investigate the origin of
the anomalous temperature dependence of the magnetocrystalline anisotropy in
(Fe1−x Cox )2 B alloys. In contrast to a conventional monotonous variation of the
MCA energy with increasing temperature, the anisotropy in (Fe1−x Cox )2 B shows a
non-monotonous temperature dependence due to increasing magnetic disorder. This
behavior of the experimental data was quantitatively reproduced by the RDLM-
based calculations. It turned out that the observed temperature dependence of the
anisotropy is caused by a modification of the electronic structure induced by spin
fluctuations which can result in a selective suppression of spin-orbit-induced hot
spots (see section “Relativistic Effects”). In contrast to the expectations based on
other models, this peculiar electronic mechanism may lead to an increase, rather
than decrease, of the anisotropy with decreasing magnetization.
Another approach to account for temperature-induced charge and spin fluctu-
ations when dealing with the magnetic properties of itinerant-electron magnets
at finite temperature is based on the functional integral method [4]. Within
this approach, the corresponding auxiliary exchange field is introduced using
the Hubbard-Stratonovich transformation [235]. This allows to describe the spin
fluctuation contribution to the free energy in a rather simple way. Based on the
functional-integral method, Kakehashi has proposed to use the dynamical coherent-
potential approximation (CPA) theory to go beyond the adiabatic theories in
metallic magnetism [236], which was first formulated for a model Hamiltonian.
In order to apply this approach to realistic systems, the dynamical CPA theory
has been extended by a combination with the LMTO band structure method
(see section “Band Structure Methods”) [237]. Using this approach, the high-
temperature susceptibility is expected to follow the Curie-Weiss law:

m2eff
χCW (T ) = . (97)
3(T − TC )

Using the atomic exchange coupling parameter J , the values J = 0.9, 0.94, and
0.9 eV for Fe, Co, and Ni and the Curie temperatures TC = 1930, 2250, and 620 K,
respectively, are found. The corresponding effective magnetic moments meff = 3.0,
expt
3.0, and 1.93 μB are in reasonable agreement with experiment (meff = 3.2 , 3.15,
and 1.6 μB , respectively)
The importance of correlation effects for finite-temperature magnetism has been
investigated within the framework of the dynamical mean field theory (DMFT) (see
section “Spin Density Functional Theory”) [16,32]. As demonstrated by Kakehashi
248 H. Ebert et al.

1 1
1000
Fe
M(T)/M(0)

χ meff 3TC

TCW (K)
2
Ni

-1
500
CPA+DMFT
Peschard 1925
Chechernikov 1962

0 0
0 0.5 1 1.5 2 0.2 0.4 0.6 0.8 1
T/TC xNi

Fig. 31 Left: temperature dependence of the magnetization M(T )/M(0) and the inverse ferro-
magnetic susceptibility for Fe (open squares) and Ni (open circles) compared with experimental
results for Fe (squares) and Ni (circles). Right: CPA+DMFT-based and experimental values for
the Curie-Weiss temperature of Fe1−x Nix alloys as a function of Ni concentration. (All data taken
from [16, 32] (left) and [239] (right))

[238], concerning the treatment of finite temperatures within electronic structure


calculations, this approach is essentially equivalent to the dynamical CPA theory
mentioned above. Figure 31 (left) shows results for the calculated temperature
dependence of the magnetic moment and the inverse ferromagnetic susceptibility of
Fe and Ni. The effective magnetic moments are found to be meff = 3.09 μB for Fe
and meff = 1.5 μB for Ni. The corresponding estimated Curie temperatures are 1900
and 700 K for Fe and Ni, respectively. A combination of the DMFT with the CPA
alloy theory that treats substitutional disorder and electronic correlations on equal
footing has been used by Poteryaev et al. [239] to investigate the magnetic properties
of Fe1−x Nix alloys. The calculated Curie temperatures shown in Fig. 31 (right) are
obviously in good agreement with experiment. Also in line with experiment, a linear
variation with temperature has been found for the calculated inverse magnetic sus-
ceptibilities at high temperatures. Recently, Patrick and Staunton have put forward a
computational scheme for the description of finite-temperature magnetic properties
of RE-TM compounds [240]. The correlation of the 4f -electrons of the RE atoms
is treated by applying the self-interaction correction (SIC) method, and the RDLM
approach is used to describe the magnetic disorder in the system. This theory was
successfully applied to the calculation of magnetic moments as a function of temper-
ature as well as the Curie temperatures of the rare-earth cobalt (RECo5 ) family for
RE = Y-Lu, demonstrating rather good agreement with experiment. Based on these
calculations, the authors proposed a mechanism responsible for the strengthening of
Re-TM as well as TM-TM interactions in the light ReCo5 compounds, where the
RE variation exhibits a strong impact on the Co-Co interactions.

Acknowledgments Financial support by the DFG through the SFB 689 and 1277 is gratefully
acknowledged.
4 Electronic Structure: Metals and Insulators 249

References
1. Mattis, D.C.: The Theory of Magnetism I, Statics and Dynamics. Springer, Berlin (1981)
2. White, R.M.: Quantum Theory of Magnetism. Springer, Berlin (2007)
3. Kübler, J.: Theory of Itinerant Electron Magnetism. International Series of Monographs on
Physics. OUP, Oxford (2009)
4. Kakehashi, Y.: Modern Theory of Magnetism in Metals and Alloys. Springer, Berlin (2012)
5. Rose, M.E.: Relativistic Electron Theory. Wiley, New York (1961)
6. Blügel, S.: Magnetische Anisotropie und Magnetostriktion (Theorie). In: 30. Ferienkurs des
Instituts für Festkörperforschung 1999 “Magnetische Schichtsysteme”, editor: Institut für
Festkörperforschung, C1.1, Forschungszentrum Jülich GmbH, Jülich (1999)
7. Chikazumi, S.: Physics of Ferromagnetism. Oxford University Press, Oxford (2009)
8. Nagaosa, N., Sinova, J., Onoda, S., MacDonald, A.H., Ong, N.P.: Anomalous Hall effect. Rev.
Mod. Phys. 82, 1539 (2010)
9. Ebert, H.: Magneto-optical effects in transition metal systems. Rep. Prog. Phys. 59, 1665
(1996)
10. Sinova, J., Valenzuela, S.O., Wunderlich, J., Back, C.H., Jungwirth, T.: Spin Hall effects. Rev.
Mod. Phys. 87, 1213 (2015)
11. Garello, K., Miron, I., Avci, C., Freimuth, F., Mokrousov, Y., Blügel, S., Auffret, S.,
Boulle, O., Gaudin, G., Gambardella, P.: Symmetry and magnitude of spin-orbit torques in
ferromagnetic heterostructures. Nat. Nanotechnol. 8, 587 (2013)
12. Heinze, S., von Bergmann, K., Menzel, M., Brede, J., Kubetzka, A., Wiesendanger, R.,
Bihlmayer, G., Blügel, S.: Spontaneous atomic-scale magnetic skyrmion lattice in two
dimensions. Nat. Phys. 7, 713 (2011)
13. Bethe, H., Salpeter, E.: Quantum Mechanics of One- and Two-Electron Atoms. Springer, New
York (1957)
14. Jansen, H.J.F.: Magnetic anisotropy in density-functional theory. Phys. Rev. B 38, 8022
(1988)
15. Engel, E., Dreizler, R.M.: Density Functional Theory – An Advanced Course. Springer, Berlin
(2011)
16. Lichtenstein, A.I., Katsnelson, M.I., Kotliar, G.: Finite-temperature magnetism of tran-
sition metals: an ab initio dynamical mean-field theory. Phys. Rev. Lett. 87, 067205
(2001)
17. Shi, J., Vignale, G., Xiao, D., Niu, Q.: Quantum Theory of Orbital Magnetization and Its
Generalization to Interacting Systems. Phys. Rev. Lett. 99, 197202 (2007)
18. Udvardi, L., Szunyogh, L., Palotás, K., Weinberger, P.: First-principles relativistic study of
spin waves in thin magnetic films. Phys. Rev. B 68, 104436 (2003)
19. Brataas, A., Tserkovnyak, Y., Bauer, G.E.W.: Scattering theory of Gilbert damping. Phys.
Rev. Lett. 101, 037207 (2008)
20. Ashcroft, N., Mermin, N.: Solid State Physics. Saunders College Publishers, New York (1976)
21. Hohenberg, P., Kohn, W.: Inhomogenous electron gas. Phys. Rev. 136, B 864 (1964)
22. Sham, L.J., Kohn, W.: One-particle properties of an inhomogeneous interacting electron gas.
Phys. Rev. 145, 561 (1966)
23. von Barth, U., Hedin, L.: A local exchange-correlation potential for the spin polarized case.
I. J. Phys. C: Solid State Phys. 5, 1629 (1972)
24. Rajagopal, A.K., Callaway, J.: Inhomogeneous electron gas. Phys. Rev. B 7, 1912 (1973)
25. Ceperley, D.M., Alder, B.J.: Ground state of the electron gas by a stochastic method. Phys.
Rev. Lett. 45, 566 (1980)
26. Ebert, H., et al.: The Munich SPR-KKR package, version 7.7, https://www.ebert.cup.uni-
muenchen.de/en/software-en/13-sprkkr (2017)
27. Leung, T.C., Chan, C.T., Harmon, B.N.: Ground-state properties of Fe, Co, Ni, and their
monoxides: results of the generalized gradient approximation. Phys. Rev. B 44, 2923 (1991)
28. Aryasetiawan, F., Gunnarsson, O.: The GW method. Rep. Prog. Phys. 61, 237
(1998)
250 H. Ebert et al.

29. Aryasetiawan, F.: Self-energy of ferromagnetic nickel in the GW approximation. Phys. Rev.
B 46, 13051 (1992)
30. Liebsch, A.: Effect of self-energy corrections on the valence-band photoemission spectra of
Ni. Phys. Rev. Lett. 43, 1431 (1979)
31. Anisimov, V.I., Zaanen, J., Andersen, O.K.: Band theory and Mott insulators: Hubbard U
instead of Stoner I. Phys. Rev. B 44, 943 (1991)
32. Kotliar, G., Savrasov, S.Y., Haule, K., Oudovenko, V.S., Parcollet, O., Marianetti, C.A.:
Electronic structure calculations with dynamical mean-field theory. Rev. Mod. Phys. 78, 865
(2006)
33. Held, K., Nekrasov, I.A., Keller, G., Eyert, V., Blümer, N., McMahan, A.K., Scalettar, R.T.,
Pruschke, T., Anisimov, V.I., Vollhardt, D.: Realistic investigations of correlated electron
systems with LDA + DMFT. Phys. Stat. Sol. (B) 243, 2599 (2006)
34. http://psi-k.net/software/
35. Blöchl, P.E.: Projector augmented-wave method. Phys. Rev. B 50, 17953 (1994)
36. Andersen, O.K.: Linear methods in band theory. Phys. Rev. B 12, 3060 (1975)
37. Singh, D.: Plane Waves, Pseudopotentials and the LAPW Method. Kluwer Academic,
Amsterdam (1994)
38. Ku, W., Berlijn, T.,Lee, C.-C.: Unfolding first-principles band structures. Phys. Rev. Lett. 104,
216401 (2010)
39. Zunger, A., Wei, S.-H., Ferreira, L.G., Bernard, J.E.: Special quasirandom structures. Phys.
Rev. Lett. 65, 353 (1990)
40. Kováčik, R., Mavropoulos, P., Wortmann, D., Blügel, S.: Spin-caloric transport properties of
cobalt nanostructures: spin disorder effects from first principles. Phys. Rev. B 89, 134417
(2014)
41. Economou, E.N.: Green’s Functions in Quantum Physics. Springer Series in Solid-State
Sciences, vol 7. Springer, Berlin (2006)
42. Ebert, H., Ködderitzsch, D., Minár, J.: Calculating condensed matter properties using the
KKR-Green’s function method – recent developments and applications. Rep. Prog. Phys. 74,
096501 (2011)
43. Soven, P.: Coherent-potential model of substitutional disordered alloys. Phys. Rev. 156, 809
(1967)
44. Butler, W.H., Stocks, G.M.: Calculated electrical conductivity and thermopower of silver-
palladium alloys. Phys. Rev. B 29, 4217 (1984)
45. Staunton, J., Gyorffy, B.L., Pindor, A.J., Stocks, G.M., Winter, H.: The ‘disordered local
moment’ picture of itinerant magnetism at finite temperatures. J. Magn. Magn. Mater. 45, 15
(1984)
46. MacDonald, A.H., Vosko, S.H.: A relativistic density functional formalism. J. Phys. C: Solid
State Phys. 12, 2977 (1979)
47. Feder, R., Rosicky, F., Ackermann, B.: Relativistic multiple scattering theory of electrons by
ferromagnets. Z. Physik B 52, 31 (1983)
48. Ebert, H.: Two ways to perform spin-polarized relativistic linear muffin-tin-orbital calcula-
tions. Phys. Rev. B 38, 9390 (1988)
49. Reiher, M., Wolf, A.: Relativistic Quantum Chemistry: The Fundamental Theory of Molecu-
lar Science. Wiley-VCH, New York (2009)
50. Pyykkö, P.: Relativistic quantum chemistry. Adv. Quantum. Chem. 11, 353 (1978)
51. Bruno, P.: Physical origins and theoretical models of magnetic anisotropy. In: Magnetismus
von Festkörpern und Grenzflächen, editor: Forschungszentrum Jülich GmbH, Institut für
Festkörperforschung, 24.1, Forschungszentrum Jülich GmbH, Jülich (1993)
52. Koelling, D.D., Harmon, B.N.: A technique for relativistic spin-polarised calculations J. Phys.
C: Solid State Phys. 10, 3107 (1977)
53. Ebert, H., Freyer, H., Vernes, A., Guo, G.-Y.: Manipulation of the spin-orbit coupling using
the Dirac equation for spin-dependent potentials. Phys. Rev. B 53, 7721 (1996)
54. Victora, R.H., MacLaren, J.M.: Predicted spin and orbital contributions to the magnetic
structure of Co/2X superlattices. J. Appl. Phys. 70, 5880 (1991)
4 Electronic Structure: Metals and Insulators 251

55. Ebert, H., Freyer, H., Deng, M.: Manipulation of the spin-orbit coupling using the Dirac
equation for spin-dependent potentials. Phys. Rev. B 56, 9454 (1997)
56. Pickel, M., Schmidt, A.B., Giesen, F., Braun, J., Minár, J., Ebert, H., Donath, M., Weinelt,
M.: Spin-orbit hybridization points in the face-centered-cubic cobalt band structure. Phys.
Rev. Lett. 101, 066402 (2008)
57. MacDonald, A.H., Daams, J.M., Vosko, S.H., Koelling, D.D.: Influence of relativistic
contributions to the effective potential on the electronic structure of Pd and Pt. Phys. Rev.
B 23, 6377 (1981)
58. Ramana, M.V., Rajagopal, A.K.: Relativistic spin-polarised electron gas. J. Phys. C: Solid
State Phys. 12, L845 (1979)
59. Ebert, H., Battocletti, M., Gross, E.K.U.: Current density functional theory of spontaneously
magnetised solids. Europhys. Lett. 40, 545 (1997)
60. Diener, G.: Current-density-functional theory for a nonrelativistic electron gas in a strong
magnetic field. J. Phys.: Cond. Mat. 3, 9417 (1991)
61. Ebert, H., Battocletti, M.: Spin and orbital polarized relativistic multiple scattering theory –
with applications to Fe, Co, Ni and Fex Co1−x . Solid State Commun. 98, 785 (1996)
62. Chadov, S., Fecher, G.H., Felser, C., Minár, J., Braun, J., Ebert, H.: Electron corre-
lations in Co2 Mn1−x Fex Si Heusler compounds. J. Phys. D: Appl. Phys. 42, 084002
(2009)
63. Chadov, S., Minár, J., Katsnelson, M.I., Ebert, H., Ködderitzsch, D., Lichtenstein, A.I.: Orbital
magnetism in transition metal systems: the role of local correlation effects. Europhys. Lett.
82, 37001 (2008)
64. Berry, M.V.: Quantal phase factors accompanying adiabatic changes. Proc. R. Soc. Lond. A:
Math. Phys. Eng. Sci. 392, 45 (1984)
65. Chang, M.-C., Niu, Q.: Berry phase, hyperorbits, and the Hofstadter spectrum: semiclassical
dynamics in magnetic Bloch bands. Phys. Rev. B 53, 7010 (1996)
66. Wu, B., Liu, J., Niu, Q.: Geometric phase for adiabatic evolutions of general quantum states.
Phys. Rev. Lett. 94, 140402 (2005)
67. Xiao, D., Chang, M.-C., Niu, Q.: Berry phase effects on electronic properties. Rev. Mod.
Phys. 82, 1959 (2010)
68. Sundaram, G., Niu, Q.: Wave-packet dynamics in slowly perturbed crystals: gradient correc-
tions and Berry-phase effects. Phys. Rev. B 59, 14915 (1999)
69. Bruno, P.: The Berry phase in magnetism and the anomalous Hall effect. In: Kronmüller, H.,
Parkin, S. (eds.) Handbook of Magnetism and Advanced Magnetic Materials, vol. 1, pp. 540–
558. Wiley, Chichester (2007)
70. Yao, Y., Kleinman, L., MacDonald, A.H., Sinova, J., Jungwirth, T., Wang, D.-S., Wang, E.,
Niu, Q.: First principles calculation of anomalous Hall conductivity in ferromagnetic bcc Fe.
Phys. Rev. Lett. 92, 037204 (2004)
71. Zhang, Y., Sun, Y., Yang, H., Železný, J., Parkin, S.P.P., Felser, C., Yan, B.: Strong anisotropic
anomalous Hall effect and spin Hall effect in the chiral antiferromagnetic compounds Mn3 X
(X = Ge, Sn, Ga, Ir, Rh, and Pt). Phys. Rev. B 95, 075128 (2017)
72. Karplus, R., Luttinger, J.M.: Hall effect in ferromagnetics. Phys. Rev. 95, 1154 (1954)
73. Jungwirth, T., Niu, Q., MacDonald, A.H.: Anomalous Hall effect in ferromagnetic semicon-
ductors. Phys. Rev. Lett. 88, 207208 (2002)
74. Berger, L.: Side-jump mechanism for the Hall effect of ferromagnets. Phys. Rev. B 2, 4559
(1970)
75. Xiao, D., Yao, Y., Fang, Z., Niu, Q.: Berry-phase effect in anomalous thermoelectric transport.
Phys. Rev. Lett. 97, 026603 (2006)
76. Xiao, D., Shi, J., Niu, Q.: Berry phase correction to electron density of states in solids. Phys.
Rev. Lett. 95, 137204 (2005)
77. Thonhauser, T., Ceresoli, D., Vanderbilt, D., Resta, R.: Orbital magnetization in periodic
insulators. Phys. Rev. Lett. 95, 137205 (2005)
78. Šmejkal, L., Jungwirth, T., Sinova, J.: Route towards Dirac and Weyl antiferromagnetic
spintronics. Phys. Status Solidi (RRL): Rapid Res. Lett. 11, 1700044 (2017)
252 H. Ebert et al.

79. Chen, H., Niu, Q., MacDonald, A.H.: Anomalous Hall effect arising from noncollinear
antiferromagnetism. Phys. Rev. Lett. 112, 017205 (2014)
80. Kübler, J., Felser, C.: Non-collinear antiferromagnets and the anomalous Hall effect. Euro-
phys. Lett. 108, 67001 (2014)
81. Šmejkal, L., Mokrousov, Y., Yan, B., MacDonald, A.H.: Topological antiferromagnetic
spintronics. Nat. Phys. 14, 242 (2018)
82. Stoner, E.C.: Collective electron specific heat and spin paramagnetism in metals. Proc. Roy.
Soc. (Lond.) A 154, 656 (1936)
83. Moriya, T.: Spin Fluctuations in Itinerant Electron Magnetism. Springer Series in Surface
Sciences, vol. 56. Springer, Berlin (1985)
84. Janak, J.F.: Uniform susceptibilities of metallic elements. Phys. Rev. B 16, 255 (1977)
85. Matsumoto, M., Staunton, J.B., Strange, P.: A new formalism for the paramagnetic
spin susceptibility of metals using relativistic spin-polarized multiple-scattering theory: a
temperature-dependent anisotropy effect. J. Phys.: Cond. Mat. 2, 8365 (1990)
86. Mankovsky, S., Ebert, H.: Theoretical description of the high-field susceptibility of magneti-
cally ordered transition metal systems with applications to Fe, Co, Ni, and Fe1−x Cox . Phys.
Rev. B 74, 54414 (2006)
87. Buczek, P., Ernst, A., Bruno, P., Sandratskii, L.M.: Energies and lifetimes of magnons in
complex ferromagnets: a first-principle study of heusler alloys. Phys. Rev. Lett. 102, 247206
(2009)
88. Şaşıoğlu, E., Schindlmayr, A., Friedrich, C., Freimuth, F., Blügel, S.: Wannier-function
approach to spin excitations in solids. Phys. Rev. B 81, 054434 (2010)
89. Gunnarsson, O.: Band model for magnetism of transition metals in the spin-density-functional
formalism. J. Phys. F: Met. Phys. 6, 587 (1976)
90. Reddy, B.V., Khanna, S.N., Dunlap, B.I.: Giant magnetic moments in 4d clusters. Phys. Rev.
Lett. 70, 3323 (1993)
91. Cox, A.J., Louderback, J.G., Apsel, S.E., Bloomfield, L.A.: Magnetism in 4d-transition metal
clusters. Phys. Rev. B 49, 12295 (1994)
92. Vondráček, M., Cornils, L., Minár, J., Warmuth, J., Michiardi, M., Piamonteze, C., Barreto,
L., Miwa, J.A., Bianchi, M., Hofmann, P., Zhou, L., Kamlapure, A., Khajetoorians, A.A.,
Wiesendanger, R., Mi, J.-L., Iversen, B.-B., Mankovsky, S., Borek, S., Ebert, H., Schüler, M.,
Wehling, T., Wiebe, J., Honolka, J.: Nickel: the time-reversal symmetry conserving partner of
iron on a chalcogenide topological insulator. Phys. Rev. B 94, 161114 (2016)
93. Dederichs, P.H., Zeller, R., Akai, H., Ebert, H.: Ab-initio calculations of the electronic
structure of impurities and alloys of ferromagnetic transition metals. J. Magn. Magn. Mater.
100, 241 (1991)
94. Hasegawa, H., Kanamori, J.: An application of the coherent potential approximation to
ferromagnetic alloys. J. Phys. Soc. Jpn. 31, 382 (1971)
95. Minár, J., Mankovsky, S., Šipr, O., Benea, D., Ebert, H.: Correlation effects in fcc-Fex Ni1x
alloys investigated by means of the KKR-CPA. J. Phys.: Cond. Mat. 26, 274206 (2014)
96. Miura, Y., Nagao, K., Shirai, M.: Atomic disorder effects on half-metallicity of the
full-Heusler alloys Co2 (Cr1−x Fex )Al: a first-principles study. Phys. Rev. B 69, 144413
(2004)
97. Galanakis, I., Mavropoulos, P., Dederichs, P.H.: Electronic structure and Slater-Pauling
behaviour in half-metallic Heusler alloys calculated from first principles. J. Phys. D: Appl.
Phys. 39, 765 (2006)
98. Galanakis, I.: Heusler Alloys. Properties, Growth, Applications. Springer Series in Material
Science, vol. 222. Springer International Publishing, Cham (2016)
99. Jourdan, M., Minár, J., Braun, J., Kronenberg, A., Chadov, S., Balke, B., Gloskovskii, A.,
Kolbe, M., Elmers, H., Schönhense, G., Ebert, H., Felser, C., Kläui, M.: Direct observation of
half-metallicity in the Heusler compound Co2 MnSi. Nat. Commun. 5, 3974 (2014)
100. Mavropoulos, P., Galanakis, I., Popescu, V., Dederichs, P.H.: The influence of spin-orbit
coupling on the band gap of Heusler alloys. J. Phys.: Cond. Mat. 16, S5759 (2004)
101. Galanakis, I.: Surface properties of the half-and full-Heusler alloys. J. Phys.: Cond. Mat. 14,
6329 (2002)
4 Electronic Structure: Metals and Insulators 253

102. Meservey, R., Tedrow, P.: Spin-polarized electron tunneling. Phys. Rep. 238, 173 (1994)
103. Mazin, I.I.: How to define and calculate the degree of spin polarization in ferromagnets. Phys.
Rev. Lett. 83, 1427 (1999)
104. Nadgorny, B.E.: Handbook of Spin Transport and Magnetism. Taylor and Francis Group,
Boca Raton (2012)
105. de Groot, R.A., Mueller, F.M., Engen, P.G.V., Buschow, K.H.J.: New class of materials: half-
metallic ferromagnets. Phys. Rev. Lett. 50, 2024 (1983)
106. Otto, M.J., van Woerden, R.A.M., van der Valk, P.J., Wijngaard, J., van Bruggen, C.F., Haas,
C., Buschow, K.H.J.: Half-metallic ferromagnets. I. Structure and magnetic properties of
NiMnSb and related inter-metallic compounds. J. Phys.: Cond. Mat. 1, 2341 (1989)
107. Galanakis, I., Dederichs, P.H., Papanikolaou, N.: Origin and properties of the gap in the half-
ferromagnetic Heusler alloys. Phys. Rev. B 66, 134428 (2002)
108. Braun, J., Ebert, H., Minár, J.: Correlation and chemical disorder in Heusler compounds: a
spectroscopical study. In: Spintronics. Fundamentals and Theory, vol. 1, Springer (2013)
109. Ishida, S., Akazawa, S., Kubo, Y., Ishida, J.: Band theory of Co2 MnSn, Co2 TiSn and Co2 TiAl.
J. Phys. F: Met. Phys. 12, 1111 (1982)
110. Fujii, S., Sugimura, S., Ishida, Asano, S.: Hyperfine fields and electronic structures of the
Heusler alloys Co2 MnX (X=Al, Ga, Si, Ge, Sn). J. Phys.: Cond. Mat. 2, 8583 (1990)
111. Galanakis, I., Dederichs, P.H., Papanikolaou, N.: Slater-Pauling behavior and origin of the
half-metallicity of the full-Heusler alloys. Phys. Rev. B 66, 174429 (2002)
112. Ozdogan, K., Galanakis, I.: First-principles electronic and magnetic properties of the half-
metallic antiferromagnet. J. Magn. Magn. Mater. 321, L34 (2009)
113. Schröter, M., Ebert, H., Akai, H., Entel, P., Hoffmann, E., Reddy, G.G.: First-principles
investigations of atomic disorder effects on magnetic and structural instabilities in transition-
metal alloys. Phys. Rev. B 52, 188 (1995)
114. van Schilfgaarde, M., Abrikosov, I.A., Johansson, B.: Origin of the Invar effect in iron-nickel
alloys. Nature 400, 46 (1999)
115. Sandratskii, L.M.: Noncollinear magnetism in itinerant-electron systems: theory and applica-
tions. Adv. Phys. 47, 91 (1998)
116. Seemann, M., Ködderitzsch, D., Wimmer, S., Ebert, H.: Symmetry-imposed shape of linear
response tensors. Phys. Rev. B 92, 155138 (2015)
117. Nakatsuji, S., Kiyohara, N., Higo, T.: Large anomalous Hall effect in a non-collinear
antiferromagnet at room temperature. Nature 527, 212 (2015)
118. Ikhlas, M., Tomita, T., Koretsune, T., Suzuki, M.-T., Nishio-Hamane, D., Arita, R., Otani, Y.,
Nakatsuji, S.: Large anomalous Nernst effect at room temperature in a chiral antiferromagnet.
Nat. Phys. 13, 1085 (2017)
119. Zhang, W., Han, W., Yang, S.-H., Sun, Y., Zhang, Y., Yan, B., Parkin, S.S.P.: Giant facet-
dependent spin-orbit torque and spin Hall conductivity in the triangular antiferromagnet
IrMn3 . Sci. Adv. 2, e1600759 (2016)
120. Železný, J., Zhang, Y., Felser, C., Yan, B.: Spin-polarized current in noncollinear antiferro-
magnets. Phys. Rev. Lett. 119, 187204 (2017)
121. Sandratskii, L.M., Kübler, J.: Magnetic structures of uranium compounds: effects of relativity
and symmetry. Phys. Rev. Lett. 75, 946 (1995)
122. Connolly, J.W.D., Williams, A.R.: Density-functional theory applied to phase transformations
in transition-metal alloys. Phys. Rev. B 27, 5169 (1983)
123. Drautz, R., Fähnle, M.: Spin-cluster expansion: parametrization of the general adiabatic
magnetic energy surface with ab initio accuracy. Phys. Rev. B 69, 104404 (2004)
124. Antal, A., Lazarovits, B., Udvardi, L., Szunyogh, L., Újfalussy, B., Weinberger, P.: First-
principles calculations of spin interactions and the magnetic ground states of Cr trimers on
Au(111). Phys. Rev. B 77, 174429 (2008)
125. Oguchi, T., Terakura, K., Hamada, N.: Magnetism of iron above the Curie temperature. J.
Phys. F: Met. Phys. 13, 145 (1983)
126. Liechtenstein, A.I., Katsnelson, M.I., Antropov, V.P., Gubanov, V.A.: Local spin density
functional approach to the theory of exchange interactions in ferromagnetic metals and alloys.
J. Magn. Magn. Mater. 67, 65 (1987)
254 H. Ebert et al.

127. Pajda, M., Kudrnovský, J., Turek, I., Drchal, V., Bruno, P.: Oscillatory Curie temperature of
two-dimensional ferromagnets. Phys. Rev. Lett. 85, 5424 (2000)
128. Polesya, S., Mankovsky, S., Šipr, O., Meindl, W., Strunk, C., Ebert, H.: Finite-temperature
magnetism of Fex Pd1−x and Cox Pt1−x alloys. Phys. Rev. B 82, 214409 (2010)
129. Ležaić, M., Mavropoulos, P., Enkovaara, J., Bihlmayer, G., Blügel, S.: Thermal collapse of
spin polarization in half-metallic ferromagnets. Phys. Rev. Lett. 97, 026404 (2006)
130. Buchelnikov, V.D., Entel, P., Taskaev, S.V., Sokolovskiy, V.V., Hucht, A., Ogura, M., Akai, H.,
Gruner, M.E., Nayak, S.K.: Monte Carlo study of the influence of antiferromagnetic exchange
interactions on the phase transitions of ferromagnetic Ni-Mn-X alloys (X=In,Sn,Sb). Phys.
Rev. B 78, 184427 (2008)
131. Buchelnikov, V.D., Sokolovskiy, V.V., Herper, H.C., Ebert, H., Gruner, M.E., Taskaev, S.V.,
Khovaylo, V.V., Hucht, A., Dannenberg, A., Ogura, M., Akai, H., Acet, M., Entel, P.:
First-principles and Monte Carlo study of magnetostructural transition and magnetocaloric
properties of Ni2+x Mn1−x Ga. Phys. Rev. B 81, 094411 (2010)
132. Sato, K., Dederichs, P.H., Katayama-Yoshida, H.: Curie temperatures of dilute magnetic
semiconductors from LDA+U electronic structure calculations. Physica B 376–377, 639
(2006)
133. Toyoda, M., Akai, H., Sato, K., Katayama-Yoshida, H.: Curie temperature of GaMnN and
GaMnAs from LDA-SIC electronic structure calculations. Phys. Stat. Sol. (C) 3, 4155 (2006)
134. Nayak, S.K., Ogura, M., Hucht, A., Akai, H., Entel, P.: Monte Carlo simulations of diluted
magnetic semiconductors using ab initio exchange parameters. J. Phys.: Cond. Mat. 21,
064238 (2009)
135. Bouzerar, G., Kudrnovský, J., Bergqvist, L., Bruno, P.: Ferromagnetism in diluted magnetic
semiconductors: a comparison between ab initio mean-field, RPA, and Monte Carlo treat-
ments. Phys. Rev. B 68, 081203 (2003)
136. Eriksson, O., Bergqvist, L., Sanyal, B., Kudrnovský, J., Drchal, V., Korzhavyi, P., Turek, I.:
Electronic structure and magnetism of diluted magnetic semiconductors. J. Phys.: Cond. Mat.
16, S5481 (2004)
137. Sato, K., Bergqvist, L., Kudmovsky, J., Dederichs, P.H., Eriksson, O., Turek, I., Sanyal,
B., Bouzerar, G., Katayama-Yoshida, H., Dinh, V.A., Fukushima, T., Kizaki, H., Zeller,
R.: First-principles theory of dilute magnetic semiconductors. Rev. Mod. Phys. 82, 1633
(2010)
138. Maccherozzi, F., Sperl, M., Panaccione, G., Minár, J., Polesya, S., Ebert, H., Wurstbauer, U.,
Hochstrasser, M., Rossi, G., Woltersdorf, G., Wegscheider, W., Back, C.H.: Evidence for a
magnetic proximity effect up to room temperature at Fe/(Ga, Mn)As interfaces. Phys. Rev.
Lett. 101, 267201 (2008)
139. Polesya, S., Šipr, O., Bornemann, S., Minár, J., Ebert, H.: Magnetic properties of free Fe
clusters at finite temperatures from first principles. Europhys. Lett. 74, 1074 (2006)
140. Šipr, O., Polesya, S., Minár, J., Ebert, H.: Influence of temperature on the systematics of
magnetic moments of free Fe clusters. J. Phys.: Cond. Mat. 19, 446205 (2007)
141. Katsnelson, M.I., Lichtenstein, A.I.: First-principles calculations of magnetic interactions in
correlated systems. Phys. Rev. B 61, 8906 (2000)
142. Ebert, H., Mankovsky, S.: Anisotropic exchange coupling in diluted magnetic semiconduc-
tors: ab initio spin-density functional theory. Phys. Rev. B 79, 045209 (2009)
143. Mankovsky, S., Bornemann, S., Minár, J., Polesya, S., Ebert, H., Staunton, J.B., Lichtenstein,
A.I.: Effects of spin-orbit coupling on the spin structure of deposited transition-metal clusters.
Phys. Rev. B 80, 014422 (2009)
144. Antropov, V.P., Katsnelson, M.I., Harmon, B.N., van Schilfgaarde, M., Kusnezov, D.: Spin
dynamics in magnets: equation of motion and finite temperature effects. Phys. Rev. B 54,
1019 (1996)
145. Ebert, H.: Relativistic theory of indirect nuclear spin-spin coupling. Phil. Mag. 88, 2673
(2008)
146. Sandratskii, L.M., Bruno, P.: Exchange interactions and Curie temperature in (Ga,Mn)As.
Phys. Rev. B 66, 134435 (2002)
4 Electronic Structure: Metals and Insulators 255

147. Uhl, M., Sandratskii, L.M., Kübler, J.: Spin fluctuations in γ -Fe and in Fe3 Pt Invar from
local-density-functional calculations. Phys. Rev. B 50, 291 (1994)
148. Heide, M., Bihlmayer, G., Blügel, S.: Dzyaloshinskii-Moriya interaction accounting for the
orientation of magnetic domains in ultrathin films: Fe/W(110). Phys. Rev. B 78, 140403
(2008)
149. Solovyev, I.V., Kashin, I.V., Mazurenko, V.V.: Mechanisms and origins of half-metallic
ferromagnetism in CrO2 . Phys. Rev. B 92, 144407 (2015)
150. Keshavarz, S., Kvashnin, Y.O., Rodrigues, D.C.M., Pereiro, M., Di Marco, I., Autieri, C.,
Nordström, L., Solovyev, I.V., Sanyal, B., Eriksson, O.: Exchange interactions of CaMnO3 in
the bulk and at the surface. Phys. Rev. B 95, 115120 (2017)
151. Logemann, R., Rudenko, A.N., Katsnelson, M.I., Kirilyuk, A.: Exchange interactions in
transition metal oxides: the role of oxygen spin polarization. J. Phys.: Condens. Matter 29,
335801 (2017)
152. Katanin, A.A., Poteryaev, A.I., Efremov, A.V., Shorikov, A.O., Skornyakov, S.L., Korotin,
M.A., Anisimov, V.I.: Orbital-selective formation of local moments in α-iron: first-principles
route to an effective model. Phys. Rev. B 81, 045117 (2010)
153. Kvashnin, Y.O., Cardias, R., Szilva, A., Di Marco, I., Katsnelson, M.I., Lichtenstein, A.I.,
Nordström, L., Klautau, A.B., Eriksson, O.: Microscopic origin of Heisenberg and Non-
Heisenberg exchange interactions in ferromagnetic bcc Fe. Phys. Rev. Lett. 116, 217202
(2016)
154. Szilva, A., Thonig, D., Bessarab, P.F., Kvashnin, Y.O., Rodrigues, D.C.M., Cardias, R.,
Pereiro, M., Nordström, L., Bergman, A., Klautau, A.B., Eriksson, O.: Theory of noncollinear
interactions beyond Heisenberg exchange: applications to bcc Fe. Phys. Rev. B 96, 144413
(2017)
155. Szunyogh, L., Újfalussy, B., Weinberger, P.: Magnetic anisotropy of iron multilayers on
Au(001): first-principles calculations in terms of the fully relativistic spin-polarized screened
KKR method. Phys. Rev. B 51, 9552 (1995)
156. Razee, S.S.A., Staunton, J.B., Pinski, F.J.: First-principles theory of magnetocrystalline
anisotropy of disordered alloys: application to cobalt platinum. Phys. Rev. B 56, 8082 (1997)
157. Újfalussy, B., Szunyogh, L., Weinberger, P.: Magnetic anisotropy in Fe/Cu(001) overlayers
and interlayers: the high-moment ferromagnetic phase. Phys. Rev. B 54, 9883 (1996)
158. Solovyev, I.V., Dederichs, P.H., Mertig, I.: Origin of orbital magnetization and magnetocrys-
talline anisotropy in TX ordered alloys (where T =Fe,Co and X =Pd,Pt). Phys. Rev. B 52,
13419 (1995)
159. Bruno, P.: Tight-binding approach to the orbital magnetic moment and magnetocrystalline
anisotropy of transition-metal monolayers. Phys. Rev. B 39, 865 (1989)
160. van der Laan, G.: Determination of the element-specific magnetic anisotropy in thin films and
surfaces. J. Phys.: Cond. Mat. 13, 11149 (2001)
161. Wang, X., Wu, R., Wang, D.-S., Freeman, A.J.: Torque method for the theoretical determina-
tion of magnetocrystalline anisotropy. Phys. Rev. B 54, 61 (1996)
162. Staunton, J.B., Szunyogh, L., Buruzs, A., Gyorffy, B.L., Ostanin, S., Udvardi, L.: Temperature
dependence of magnetic anisotropy: an ab initio approach. Phys. Rev. B 74, 144411 (2006)
163. Daalderop, G.H.O., Kelly, P.J., Schuurmans, M.F.H.: First-principles calculation of the
magnetic anisotropy energy of (Co)n /(X)m multilayers. Phys. Rev. B 42, 7270 (1990)
164. Weinberger, P.: Magnetic Anisotropies in Nanostructured Matter. Condensed Matter Physics.
Chapman and Hall/CRC Press, Boca Raton (2008)
165. Stiles, M.D., Halilov, S.V., Hyman, R.A., Zangwill, A.: Spin-other-orbit interaction and
magnetocrystalline anisotropy. Phys. Rev. B 64, 104430 (2001)
166. Bornemann, S., Minár, J., Braun, J., Ködderitzsch, D., Ebert, H.: Ab-initio description of the
magnetic shape anisotropy due to the Breit interaction. Solid State Commun. 152, 85 (2012)
167. Buschow, K., van Diepen, A., de Wijn, H.: Crystal-field anisotropy of Sm3+ in SmCo5 . Solid
State Commun. 15, 903 (1974)
168. Yamada, M., Kato, H., Yamamoto, H., Nakagawa, Y.: Crystal-field analysis of the magneti-
zation process in a series of Nd2 Fe14 B-type compounds. Phys. Rev. B 38, 620 (1988)
256 H. Ebert et al.

169. Herbst, J.F.: R2 Fe14 B materials: intrinsic properties and technological aspects. Rev. Mod.
Phys. 63, 819 (1991)
170. Hummler, K., Fähnle, M.: Full-potential linear-muffin-tin-orbital calculations of the magnetic
properties of rare-earth transition-metal intermetallics. I. Description of the formalism and
application to the series R Co5 (R =rare-earth atom). Phys. Rev. B 53, 3272 (1996)
171. Hummler, K., Fähnle, M.: Ab initio calculation of local magnetic moments and the crystal
field in R2 Fe14 B (R =Gd, Tb, Dy, Ho, and Er). Phys. Rev. B 45, 3161 (1992)
172. Coehoorn, R.: Supermagnets, Hard Magnetic Materials. Nato ASI Series, Series C, chapter 8,
vol. 331, p. 133. Kluwer Academic Publishers, Dardrecht (1991)
173. Richter, M., Oppeneer, P.M., Eschrig, H., Johansson, B.: Calculated crystal-field parameters
of SmCo5 . Phys. Rev. B 46, 13919 (1992)
174. Hummler, K., Fähnle, M.: Full-potential linear-muffin-tin-orbital calculations of the magnetic
properties of rare-earth transition-metal intermetallics. II. Nd2 Fe14 B. Phys. Rev. B 53, 3290
(1996)
175. Moriya, H., Tsuchiura, H., Sakuma, A.: First principles calculation of crystal field parameter
near surfaces of Nd2 Fe14 B. J. Appl. Phys. 105, 07A740 (2009)
176. Tanaka, S., Moriya, H., Tsuchiura, H., Sakuma, A., Diviš, M., Novák, P.: First principles study
on the local magnetic anisotropy near surfaces of Dy2 Fe14 B and Nd2 Fe14 B magnets. J. Appl.
Phys. 109, 07A702 (2011)
177. Novák, P., Knížek, K., Kuneš, J.: Crystal field parameters with Wannier functions: application
to rare-earth aluminates. Phys. Rev. B 87, 205139 (2013)
178. Novák, P., Kuneš, J., Knížek, K.: Crystal field of rare earth impurities in LaF3 . Opt. Mater.
37, 414 (2014)
179. Patrick, C.E., Kumar, S., Balakrishnan, G., Edwards, R.S., Lees, M.R., Petit, L., Staunton,
J.B.: Calculating the magnetic anisotropy of rare-earth–transition-metal ferrimagnets. Phys.
Rev. Lett. 120, 097202 (2018)
180. Halilov, S.V., Eschrig, H., Perlov, A.Y., Oppeneer, P.M.: Adiabatic spin dynamics from spin-
density-functional theory: application to Fe, Co, and Ni. Phys. Rev. B 58, 293 (1998)
181. Grotheer, O., Ederer, C., Fähnle, M.: Fast ab initio methods for the calculation of adiabatic
spin wave spectra in complex systems. Phys. Rev. B 63, 100401 (2001)
182. Pajda, M., Kudrnovský, J., Turek, I., Drchal, V., Bruno, P.: Ab initio calculations of exchange
interactions, spin-wave stiffness constants, and Curie temperatures of Fe, Co, and Ni. Phys.
Rev. B 64, 174402 (2001)
183. Turek, I., Kudrnovský, J., Drchal, V., Bruno, P.: Exchange interactions, spin waves, and
transition temperatures in itinerant magnets. Phil. Mag. 86, 1713 (2006)
184. Brinkman, W.F., Elliot, R.J.: Theory of spin-space groups. Proc. R. Soc. (Lond.) A 294, 343
(1966)
185. Brinkman, W.F., Elliot, R.J.: Space group theory for spin waves. J. Appl. Phys. 37, 1457
(1966)
186. Herring, C.: Magnetism: exchange interactions among itinerant electrons In: Rado, G., Suhl,
H. (eds.) Magnetism, vol. IV, p. 191. Academic Press, New York (1966)
187. Sandratskii, L.M.: Symmetry analysis of electronic states for crystals with spiral magnetic
order. I. General properties. J. Phys.: Cond. Mat. 3, 8565 (1991)
188. Uhl, M., Sandratskii, L., Kübler, J.: Electronic and magnetic states of γ -Fe. J. Magn. Magn.
Mater. 103, 314 (1992)
189. Kurz, P., Förster, F., Nordström, L., Bihlmayer, G., Blügel, S.: Ab initio treatment of
noncollinear magnets with the full-potential linearized augmented plane wave method. Phys.
Rev. B 69, 024415 (2004)
190. Mankovsky, S., Fecher, G.H., Ebert, H.: Electronic structure calculations in ordered and
disordered solids with spiral magnetic order. Phys. Rev. B 83, 144401 (2011)
191. Kübler,J.: Ab initio estimates of the Curie temperature for magnetic compounds. J. Phys.:
Condens. Matter 18, 9795 (2006)
192. Galanakis, I., Sasioglu, E.: Ab-initio calculation of effective exchange interactions, spin
waves, and Curie temperature in L21 - and L12 -type local moment ferromagnets. J. Mater.
Sci. 47, 7678 (2012)
4 Electronic Structure: Metals and Insulators 257

193. Şaşıoğlu, E., Sandratskii, L.M., Bruno, P., Galanakis, I.: Exchange interactions and temper-
ature dependence of magnetization in half-metallic Heusler alloys. Phys. Rev. B 72, 184415
(2005)
194. Edwards, D.M., Katsnelson, M.I.: High-temperature ferromagnetism of sp electrons in narrow
impurity bands: application to CaB6 . J. Phys.: Condens. Matter 18, 7209 (2006)
195. Buczek, P., Ernst, A., Sandratskii, L.M.: Spin dynamics of half-metallic Co2 MnSi. J. Phys.:
Conf. Ser. 200, 042006 (2010)
196. Savrasov, S.Y.: Linear response calculations of spin fluctuations. Phys. Rev. Lett. 81, 2570
(1998)
197. Qian, Z., Vignale, G.: Spin dynamics from time-dependent spin-density-functional theory.
Phys. Rev. Lett. 88, 056404 (2002)
198. Lounis, S., dos Santos Dias, M., Schweflinghaus, B.: Transverse dynamical magnetic
susceptibilities from regular static density functional theory: evaluation of damping and g
shifts of spin excitations. Phys. Rev. B 91, 104420 (2015)
199. Bruno, P.: Exchange interaction parameters and adiabatic spin-wave spectra of ferromagnets:
a “renormalized magnetic force theorem”. Phys. Rev. Lett. 90, 087205 (2003)
200. Katsnelson, M.I., Lichtenstein, A.I.: Magnetic susceptibility, exchange interactions and spin-
wave spectra in the local spin density approximation. J. Phys.: Condens. Matter 16, 7439
(2004)
201. Buczek, P., Ernst, A., Sandratskii, L.M.: Interface electronic complexes and landau damping
of magnons in ultrathin magnets. Phys. Rev. Lett. 106, 157204 (2011)
202. Tajima, K., Ishikawa, Y., Webster, P.J., Stringfellow, M.W., Tocchetti, D., Zeabeck, K.R.A.:
Spin waves in a heusler alloy Cu2 MnAl. J. Phys. Soc. Jpn. 43, 483 (1977)
203. Buczek, P., Ernst, A., Sandratskii, L.M.: Different dimensionality trends in the Landau
damping of magnons in iron, cobalt, and nickel: time-dependent density functional study.
Phys. Rev. B 84, 174418 (2011)
204. Müller, M.C.T.D., Friedrich, C., Blügel, S.: Acoustic magnons in the long-wavelength limit:
investigating the Goldstone violation in many-body perturbation theory. Phys. Rev. B 94,
064433 (2016)
205. Staunton, J.B., Poulter, J., Ginatempo, B., Bruno, E., Johnson, D.D.: Incommensurate
and commensurate antiferromagnetic spin fluctuations in Cr and Cr alloys from ab initio
dynamical spin susceptibility calculations. Phys. Rev. Lett. 82, 3340 (1999)
206. Staunton, J.B., Poulter, J., Ginatempo, B., Bruno, E., Johnson, D.D.: Spin fluctuations in
nearly magnetic metals from ab initio dynamical spin susceptibility calculations: application
to Pd and Cr95 V5 . Phys. Rev. B 62, 1075 (2000)
207. Schindlmayr, A., Friedrich, C., Sasioglu, E., Blügel, S.: First-principles calculation of
electronic excitations in solids with SPEX. Z. Phys. Chem. 224, 357 (2010)
208. Tyablokov, S.V.: Methods of Quantum Theory of Magnetism. Plenum Press, New York
(1967)
209. Callen, H.B.: Green function theory of ferromagnetism. Phys. Rev. 130, 890 (1963)
210. Bose, S.K., Kudrnovský, J.,Drchal, V., Turek, I.: Magnetism of mixed quaternary Heusler
alloys: (N i, T )2 MnSn (T = Cu, P d) as a case study. Phys. Rev. B 82, 174402 (2010)
211. Rusz, J., Turek, I., Diviš, M.: Random-phase approximation for critical temperatures of
collinear magnets with multiple sublattices: GdX compounds (X = Mg, Rh, Ni, Pd). Phys.
Rev. B 71, 174408 (2005)
212. Sandratskii, L.M., Singer, R., Şaşıoğlu, E.: Heisenberg Hamiltonian description of multiple-
sublattice itinerant-electron systems: general considerations and applications to NiMnSb and
MnAs. Phys. Rev. B 76, 184406 (2007)
213. Mermin, N.D., Wagner, H.: Absence of ferromagnetism or antiferromagnetism in one- or
two-dimensional isotropic Heisenberg models. Phys. Rev. Lett. 17, 1133 (1966)
214. Bruno, P.: Magnetization and Curie temperature of ferromagnetic ultrathin films: the influence
of magnetic anisotropy and dipolar interactions (invited). Mater. Res. Soc. Symp. Proc. 231,
299 (1991)
215. Bouzerar, G., Bruno, P.: RPA-CPA theory for magnetism in disordered Heisenberg binary
systems with long-range exchange integrals. Phys. Rev. B 66, 014410 (2002)
258 H. Ebert et al.

216. Kübler, J., Fecher, G.H., Felser, C.: Understanding the trend in the Curie temperatures of
Co2 -based Heusler compounds: ab initio calculations. Phys. Rev. B 76, 024414 (2007)
217. Rosengaard, N.M., Johansson, B.: Finite-temperature study of itinerant ferromagnetism in Fe,
Co, and Ni. Phys. Rev. B 55, 14975 (1997)
218. Jakobsson, A., Şaşıoğlu, E., Mavropoulos, P., Ležaić, M., Sanyal, B., Bihlmayer, G., Blügel,
S.: Tuning the Curie temperature of FeCo compounds by tetragonal distortion. Appl. Phys.
Lett. 103, 102404 (2013)
219. Bergqvist, L., Korzhavyi, P.A., Sanyal, B., Mirbt, S., Abrikosov, I.A., Nordström, L.,
Smirnova, E.A., Mohn, P., Svedlindh, P., Eriksson, O.: Magnetic and electronic structure of
(Ga1−x Mnx )As. Phys. Rev. B 67, 205201 (2003)
220. Bergqvist, L., Eriksson, O., Kudrnovský, J., Drchal, V., Korzhavyi, P., Turek, I.: Magnetic
percolation in diluted magnetic semiconductors. Phys. Rev. Lett. 93, 137202 (2004)
221. Ležaić, M., Mavropoulos, P., Bihlmayer, G., Blügel, S.: Exchange interactions and local-
moment fluctuation corrections in ferromagnets at finite temperatures based on noncollinear
density-functional calculations. Phys. Rev. B 88, 134403 (2013)
222. Uhl, M., Kübler, J.: Exchange-coupled spin-fluctuation theory: application to Fe, Co, and Ni.
Phys. Rev. Lett. 77, 334 (1996)
223. Ruban, A.V., Khmelevskyi, S., Mohn, P., Johansson, B.: Temperature-induced longitudinal
spin fluctuations in Fe and Ni. Phys. Rev. B 75, 054402 (2007)
224. Williams, A.R., Zeller, R., Moruzzi, V.L., Gelatt, C.D., Kubler, J.: Covalent magnetism: an
alternative to the Stoner model. J. Appl. Phys. 52, 2067 (1981)
225. Mohn, P., Schwarz, K.: Supercell calculations for transition metal impurities in palladium. J.
Phys.: Cond. Mat. 5, 5099 (1993)
226. Mryasov, O.N., Nowak, U., Guslienko, K.Y., Chantrell, R.W.: Temperature-dependent mag-
netic properties of FePt: effective spin Hamiltonian model. Europhys. Lett. 69, 805 (2005)
227. Mryasov, O.N.: Magnetic interactions and phase transformations in FeM, M = (Pt,Rh) ordered
alloys. Phase Transit. 78, 197 (2005)
228. Kudrnovský, J., Drchal, V., Bruno, P.: Magnetic properties of fcc Ni-based transition metal
alloys. Phys. Rev. B 77, 224422 (2008)
229. Polesya, S., Mankovsky, S., Ködderitzsch, D., Minár, J., Ebert, H.: Finite-temperature
magnetism of FeRh compounds. Phys. Rev. B 93, 024423 (2016)
230. Gyorffy, B.L., Pindor, A.J., Staunton, J., Stocks, G.M., Winter, H.: A first-principles theory
of ferromagnetic phase transitions in metals. J. Phys. F: Met. Phys. 15, 1337 (1985)
231. Feynman, R.P.: Slow electrons in a polar crystal. Phys. Rev. 97, 660 (1955)
232. Staunton, J.B., Ostanin, S., Razee, S.S.A., Gyorffy, B.L., Szunyogh, L., Ginatempo, B.,
Bruno, E.: Temperature dependent magnetic anisotropy in metallic magnets from an ab initio
electronic structure theory: L10 -ordered FePt. Phys. Rev. Lett. 93, 257204 (2004)
233. Buruzs, A., Weinberger, P., Szunyogh, L., Udvardi, L., Chleboun, P.I., Fischer, A.M.,
Staunton, J.B.: Ab initio theory of temperature dependence of magnetic anisotropy in layered
systems: applications to thin Co films on Cu(100). Phys. Rev. B 76, 064417 (2007)
234. Zhuravlev, I.A., Antropov, V.P., Belashchenko, K.D.: Spin-fluctuation mechanism of anoma-
lous temperature dependence of magnetocrystalline anisotropy in itinerant magnets. Phys.
Rev. Lett. 115, 217201 (2015)
235. Hubbard, J.: Calculation of partition functions. Phys. Rev. Lett. 3, 77 (1959)
236. Kakehashi, Y.: Monte Carlo approach to the dynamical coherent-potential approximation in
metallic magnetism. Phys. Rev. B 45, 7196 (1992)
237. Kakehashi, Y., Shimabukuro, T., Tamashiro, T., Nakamura, T.: Dynamical coherent-potential
approximation and tight-binding linear muffintin orbital approach to correlated electron
system. J. Phys. Soc. Jpn. 77, 094706 (2008)
238. Kakehashi, Y.: Many-body coherent potential approximation, dynamical coherent potential
approximation, and dynamical mean-field theory. Phys. Rev. B 66, 104428 (2002)
239. Poteryaev, A.I., Skorikov, N.A., Anisimov, V.I., Korotin, M.A.: Magnetic properties of
Fe1−x Nix alloy from CPA+DMFT perspectives. Phys. Rev. B 93, 205135 (2016)
4 Electronic Structure: Metals and Insulators 259

240. Patrick, C.E., Staunton, J.B.: Rare-earth/transition-metal magnets at finite temperature: self-
interaction-corrected relativistic density functional theory in the disordered local moment
picture Phys. Rev. B 97, 224415 (2018)

Hubert Ebert studied physics and received his Ph.D. from the
Ludwig-Maximilians-University Munich in 1986. After a post-
doc stay at the University of Bristol (UK), he worked for several
years at the central laboratory for research and development of
Siemens Company in Erlangen. Since 1993 he is professor for
theoretical physical chemistry at the university of Munich.

Sergiy Mankovsky studied physics in Moscow institute of


Physics and Technology. In 1992 he received the degree of candi-
date of physico-mathematical sciences (an analogue of the Ph.D.)
from the Kurdyumov Institute for Metal Physics of the N.A.S. of
Ukraine, where he worked as a research scientist until 2001. Since
2001 he works at the Ludwig-Maximilians-University Munich.

Sebastian Wimmer received his Ph.D. from the Ludwig-


Maximilians-Universität München in 2018. Until 2019 he worked
in the group of Prof. Dr. Hubert Ebert, focusing on the first-
principles description of spintronic and spincaloritronic linear
response properties of metals and alloys.
Quantum Magnetism
5
Gabriel Aeppli and Philip Stamp

Contents
Spin Paths and Spin Phase . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 262
The Importance of Decoherence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 265
Quantum Relaxation in Dipolar Nets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 268
Large-Scale Coherence and Entanglement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 271
Future Directions and Open Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 276
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 277

Abstract

Macroscopic quantum effects have been familiar since the discovery of super-
fluids and superconductors over 100 years ago. In the last few decades, it
has been understood how large-scale quantum effects can also show up in
“spin space.” The collective tunneling of many spins was observed in magnetic
nanomolecules and in insulating dipolar-coupled spin arrays, and the tunneling
of ferromagnetic domain walls has also been cleanly identified. To see large-
scale coherence or entanglement effects, the decoherence caused by interactions
with the environment (particularly with nuclear spins) must be controlled.
Theory indicates ways of doing this, and systems ranging from classic magnetic
compounds to deterministically doped silicon will make the job easier. Coherent

G. Aeppli ()
Physics Department (ETHZ), Institut de Physique (EPFL) and Photon Science Division (PSI),
ETHZ, EPFL and PSI, Zürich, Lausanne and Villigen, Switzerland
e-mail: aepplig@ethz.ch
P. Stamp ()
Pacific Institute of Theoretical Physics, University of British Columbia, Vancouver, BC, Canada
e-mail: stamp@phas.ubc.ca

© Springer Nature Switzerland AG 2021 261


J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_5
262 G. Aeppli and P. Stamp

control of quantum spin arrays, and large-scale quantum spin superpositions, is


a likely prospect for the future.

Magnetism has its microscopic origin in quantum mechanics which determines


basic parameters such as single-site anisotropy and intersite couplings. However,
once the parameters have been fixed, classical reasoning is usually all that is
required to model and engineer real materials and devices. In this sense, magnetism
is not different than any other branch of condensed matter physics: quantum
mechanics is essential to describe very small objects such as individual nuclei,
atoms, and molecules, but the emergent behavior of macroscopic ensembles of
such objects can almost always be explained in classical terms. There are a few
striking exceptions, most notably for fractional quantum Hall systems and Bose
condensates. These depend for their existence on “macroscopic wave functions”
[1], which correlate simultaneously the dynamics of all the particles into a single-
wave function Ψ ∼ |Ψ (r, t)|eiϕ(r,t) , with a phase ϕ(r, t) varying in space and time.
So, can any other kinds of large-scale quantum behavior be expected in physics, not
entailing a Laughlin state, Bose condensation, or supercurrents? We report here on
some of the remarkable ways this can happen, not in real space but in spin space.
These developments are relatively recent and constitute a new class of quantum
phenomena – and in the last three decades or so, experiments have begun to confront
theory in a rigorous way.

Spin Paths and Spin Phase

There is a remarkable way of formulating quantum mechanics, found by Feynman


during his doctoral studies [2], which allows a neat appreciation of these develop-
ments. One writes the usual amplitude Gba , for the transition between two quantum
states |ψa  and |ψb , as a sum of “amplitudes” over all possible paths between
them – the system simultaneously explores all these paths. The probability that
the transition will then occur is the usual “amplitude squared,” i.e., |Gba |2 . The
amplitude or weighting factor for the μ-th path is ei Aμ /h̄ , exponentiating a phase
ϕ = Aμ /h̄, where Aμ is the “action” of this path. Crucially, the action is just that
for the corresponding
 classical system (e.g., a particle) moving along the same path,
i.e., Aμ = dtL, where L is the classical Lagrangian.
This sum over paths, or “path integral,” brings out vividly the essential role of
phase interference between different paths. It has been enormously fruitful, both
as a pedagogical guide [3] and in advancing our understanding of basic quantum
physics and quantum field theory [4]. An obvious question, which notoriously
worried Feynman, is how to then deal with quantum spins, which have no classical
limit (if the spin quantum number S is finite, then when h̄ → 0, the spin moment
h̄S disappears!). The answer turned out to be interesting [5, 6, 7]. Imagine a unit
sphere in spin space, with a “particle” of unit charge moving on its surface – its
coordinate n̂(t) = S(t)/S defines the spin direction (see Fig. 1). The fake charge
5 Quantum Magnetism 263

accumulates a conventional “potential phase” by coupling to a potential Ho (S) (the


spin Hamiltonian) on the sphere. Crucially, however, the charge also couples to a
fake monopole of strength q = h̄S, situated at the center of the sphere. This adds a
“kinetic” or Berry phase [6], having the same form as the Aharonov-Bohm phase [8]
for a charge moving in real space (see Fig. 1a). One then sums over paths on the spin
sphere, to reproduce exactly the dynamics of a quantum spin.
Tunneling spins: As an example, suppose Ho (S) has local potential minima – at
low energy, the spin must then quantum tunnel between these minima to move at
all. For simplicity, we imagine two paths, connecting the lowest states | ↑, | ↓ in
these energy minima (Fig. 1b). This is just like the famous “two-slit” problem – the
tunneling amplitude o will be controlled by the interference due to the different
phases ϕ1 , ϕ2 accumulated along the two paths [3]. Two points are interesting here.
First, as discussed by Bogachek and Krive [10], and later others [11, 12], one can
manipulate the two paths by applying a transverse magnetic field H⊥ o , thereby giving
oscillations in the tunneling amplitude (Fig. 1c). Second, by considering only the
lowest state in each potential well, we have in effect truncated the Hamiltonian
Ho (S) to a two-level system (justified if the temperature is low enough that higher
spin states are inactive). We can write Ho (S) → Ho (τ̂ ) = o τ̂x , where τ̂ is the
Pauli vector. The eigenstates are just the symmetric and antisymmetric combinations
of | ↑ and | ↓. If the paths are oppositely directed but otherwise symmetric, then
ϕ1 = −ϕ2 = π S, and the transition amplitude (and hence the tunneling splitting)
between the two eigenstates is then ∼ (eiϕ +e−iϕ ) ∝ cos(π S), giving a gap between
the lowest and first excited states if S is integer-valued, but no gap if S is half-integer.
This last result was of course already noted by Kramers [13] in 1930; spin paths are
not required to understand it! Interest in spin phases was really set off by Haldane’s
remarkable prediction [9] that a large gap existed in the spectrum of an integer spin
chain (but not a half-integer one).
If we add an extra “longitudinal” field Hoz to lift the degeneracy of the minima
(i.e., of the two states | ↑ and | ↓), by an “energy bias,” o = gμB Sz Hoz . Then our
two-level system (or qubit, in the language of quantum information) has
Hamiltonian:

Ho (τ̂ ) = o τ̂ x + o τ̂ z (1)

We see the tunneling term o tries to drive quantum transitions between | ↑ and
| ↓, but is hindered by the longitudinal field bias o , which pushes the states | ↑,
| ↓ out of resonance. Notice that we can manipulate both o and o with external
fields.
Quantum Ising Spin Networks: Now imagine an interacting “spin net” of two-
level systems τ̂i , with i = 1, 2, ..N, and each with its own tunneling amplitude i
and bias i . Adding interspin interactions between them generalizes (1) to:
  
HoQI = i τ̂ix + i τ̂iz + Vij τ̂iz τ̂jz (2)
i ij
264 G. Aeppli and P. Stamp

a b
flux

ω n(t)

c 1E+1

1E+0

1E-1

1E-2
o
1E-3 90

1E-4 o
50
Δ(K)
1E-5
o
20
1E-6
o
1E-7 7
o
1E-8 0

1E-9

1E-10
0 1 2 3 4 5
H (T)

Fig. 1 (Continued)
5 Quantum Magnetism 265

We only include longitudinal couplings (coupling the z-components of the spins) in


the interaction Vij ; in many interesting magnetic systems, the other interaction terms
are strongly suppressed relative to these. Notice that the total longitudinal
 field bias
blocking quantum transitions on the i-th site is now ξi = i + j Vij τjz , instead of
just i .
This “Quantum Ising” model describes a large variety of physical systems which,
at low temperature, truncate to a set of interacting two-level systems. The best
known examples in quantum magnetism are dipolar-coupled magnetic molecules
and ions and quantum spin glasses. However, H0QI also describes a set of N
interacting qubits – and if we can manipulate the couplings in it, we have a toy
model for a quantum computer. In quantum computation (reviewed in [14, 15, 16]),
one creates and manipulates N-qubit states which are “entangled,” i.e., which cannot
in general be written as products over separate spins. Such states are fundamentally
nonclassical, as first discussed by Schrödinger and Einstein in 1935. The “quantum
information” in them is encoded in the 2N −1 relative “spin phases” between the
different qubits. To make and use such states is one of the great goals in this field –
but it will be hard. To see why, we must first understand the main obstacle in the
way.

The Importance of Decoherence

Decoherence arises when a quantum system interacts with its environment, and their
Feynman paths and quantum phases become entangled. Even if they later decouple,
averaging over the unknown environmental states then “smears” over states of
the system, destroying phase coherence over some timescale τφ (the “decoherence
time”). Decoherence in many-particle systems is usually much larger than expected,


Fig. 1 Path integrals on the spin sphere. In (a), we show a spin S as a unit charge, moving on the
unit sphere around a magnetic monopole of strength q = h̄S, along a path which  has coordinate
n̂(t) = S(t)/S, and encloses a solid angle . The kinetic phase φB = q/h̄ A · dn, where A
is the monopole vector potential (compare the “Aharonov-Bohm” phase [8] accumulated by a
charged particle moving through an ordinary magnetic vector potential A(r) in real space). For
a closed path, Gauss’s theorem then shows that the enclosed flux from the fake monopole is just
φB = S. In (b), a biaxial (easy ẑ-axis, hard x̂-axis) potential field Ho (S) is added (dark areas
are regions of higher potential). The spin moves preferentially between the two minimum energy
states at the poles by tunneling along the pair of minimum action paths (shown as dashed lines),
with amplitudes 12 μ eiϕμ respectively, where μ = 1, 2 labels the paths, and the μ are real.
An external field H⊥ o , applied along the hard x̂-axis, pulls the two states, and the paths between
them, toward x̂, thereby reducing the enclosed area on the unit sphere. In this “symmetric”
case, ϕ1 = −ϕ2 = ϕ(H⊥ ˜
o ) and 1 = 2 = o . The total tunneling amplitude  is then just the
sum 12 o (eiϕ + e−iϕ ), i.e.,  ˜ o (H⊥
o ) = o (H ⊥ ) cos ϕ(H⊥ ). (c) shows the resulting oscillations in
o o
˜ o (H⊥o ), for a typical biaxial potential (easy axis ẑ, hard axis x̂), as a function of transverse field
H⊥ ⊥
o . If Ho is rotated away from x̂ by an angle φ (shown here in degrees), then |ϕ1 | = |ϕ2 |, and
1 = 2 , i.e., one path is favored over the other, and the oscillations are lost
266 G. Aeppli and P. Stamp

a b

+ E(t)

ε(t)
t=0

γk
γk
- E(t)

c Vo Dipolar-Dominated
Regime

Li Ho x Y1-x F4
0.2 K

Mn-12 Quantum

0.1 K
ξ o Relaxation
Regime

Fe-8
1K
Δo Quantum
Coherence
Regime

Fig. 2 (Continued)
5 Quantum Magnetism 267

and still somewhat mysterious (witness the debate over the saturation of dephasing
times in mesoscopic conductors [17, 18]). Moreover, superpositions and entangled
states are extraordinarily sensitive to even very small environmental interactions.
To get a feeling for decoherence, let us go back to our toy tunneling spin, and now
couple it to a bath of “satellite” spins. In the real world, these satellite spins describe
localized modes (defects, nuclear and paramagnetic spins, etc.) which couple to the
central spin [19]. There are also delocalized environmental modes like electrons,
photons, phonons, etc., which also cause decoherence [21, 22], and which can be
described as a bath of oscillators [20]).
How do the satellite spins and oscillators dephase the “central” tunneling spin?
We explain this again using path integral language (Fig. 2). There are three main
decoherence mechanisms:

(i) Typically the environmental modes have their own dynamics, creating an
extra fluctuating “noise” field on the central system. This adds random phases
to each path, eventually destroying phase coherence between them (“noise
decoherence”). The noise can even push the central spin in and out of resonance
(Fig. 2a).
(ii) When the central spin tunnels, it causes a sudden “kick” perturbation on
the satellite spins, giving them an extra phase which is entangled with the
central spin phase – thence dephasing the central spin dynamics (“topological
decoherence” [23]).
(iii) The field on the k-th satellite, from the central spin, flips with the central
spin between two (in general noncollinear) orientations (Fig. 2b). Between
flips, the satellite precesses in these fields. Summing over all central spin
paths, each involving a different accumulated satellite precessional phase,


Fig. 2 The role of decoherence: In (a), we see the effect of a randomly fluctuating environmental
noise bias ε(t) (black curve) on a tunneling two-level qubit with tunneling matrix element o .
The two levels having adiabatic energies ±E(t), with E 2 (t) = 2o + ε2 (t), are shown as red
and blue curves. The system can only make transitions when near “resonance” (i.e., when |ε(t)|
is ∼ o or less, the regions shown in green). In (b), we show schematically the motion of a
satellite spin, in the presence of a qubit which is flipping between two different states | ↑ and
| ↓. When the qubit flips, the qubit field acting on the k-th satellite spin rapidly changes, from
↑ ↓
γk to γk (or vice versa). Between flips, the spin precesses around the qubit field, accumulating
an extra “precessional” phase. Averaging over this phase gives precessional decoherence. The
sudden change of qubit field also perturbs the satellite spin phase, giving further decoherence
(the “topological decoherence” mechanism [23]). (c) shows the important parameters for a spin
network – the typical tunneling splitting o , the characteristic energy Vo of interspin interactions,
and the energy scale ξo governing interactions with the environment (which in insulating magnetic
systems at very low T comes from the coupling to nuclear spins). For definiteness, we show the
parameter range covered in this space by experiments in crystals of Mn-12 and F e-8 molecular
magnets, and in LiH ox Y1−x F4 ; energy scales are in temperature units. In these systems, o is
controlled by varying the transverse field, and Vo is varied by changing x (in LiH ox Y1−x F4 ), or
by diluting the molecules in solution (in F e-8 and Mn-12)
268 G. Aeppli and P. Stamp

gives “precessional” decoherence. In an oscillator bath, the central spin flip


slightly shifts the oscillator wave functions – for metallic environments, this
gives Anderson’s “orthogonality catastrophe” [24], a very strong decoherence
mechanism.

Notice that decoherence may involve very little energy transfer – it is not
necessarily a dissipative process. In magnetic systems, the worst low-T decoherence
will come from very low-energy localized modes, particularly nuclear spins,
which cause very little dissipation, but lots of precessional decoherence [19, 25].
Delocalized modes like phonons, photons, and electrons cause strong decoherence
(and strong dissipation) at higher energy, where they have a high density of states.
Thus, at intermediate energies (typically around 0.01 − 0.5 K), decoherence is at a
minimum. This “window” of low decoherence is of great practical importance – it
also exists for many other solid-state systems [26].
The basic problem with our toy model (2) for a quantum computer is now
clear – it ignores decoherence. If we couple each of the spins in the spin net to
an environment, there are now three main energy scales (Fig. 2c). A “quantum”
parameter o (the typical value of i ) drives the dynamics, along with interspin
interactions of typical strength Vo ; but a coupling of each spin to the environment,
having some effective energy scale ξo , destroys phase coherence. If we could switch
off ξo (i.e., stay in the Vo − o plane in Fig. 2c), we would have perfect quantum
behavior, with Vij correlating the entangled dynamics of vast numbers of qubits –
this would be true macroscopic quantum spin entanglement. But how close are we
to this goal?

Quantum Relaxation in Dipolar Nets

In fact most work has been done on systems with dipolar interspin couplings, having
non-negligible environmental interactions – i.e., near the Vo − ξo plane in Fig. 2c.
These systems are very complex – but a simple theoretical picture can be given.
Consider first a single spin qubit τi . The net bias i on τiz now includes a dynamic
contribution from the nuclear spin environment. This typically fluctuates over an
energy range ξo ∼ Eo , where Eo defines the energy width of the multiplet of nuclear
spin states coupled to S; this width is easy to calculate if the hyperfine couplings are
known. Then if i is within a “tunneling window” of width ξo around zero bias, the
fluctuating field can actually bring the qubit to resonance (recall Fig. 2a), where it
can make inelastic (i.e., incoherent) transitions [25].
Now consider an interacting network – assuming here for definiteness that go =
o /Vo
1 (the “dipolar-dominated” regime in Fig. 2c). As the resonant spins
tunnel, a “hole” of width ξo should appear in the distribution of longitudinal
 fields
in the system, around zero (see Fig. 3a). The interaction contribution j Vij τj to i
then plays a key role – it slowly changes as the τj relax, bringing more spins into
the tunneling window (hole “refilling”). The total spin distribution is then predicted
5 Quantum Magnetism 269

a [P (ξ,t) - P ( ξ,t)]
εo

Vo

2ξo

0 ξ

Fig. 3 (Continued)
270 G. Aeppli and P. Stamp

Fig. 3 Collective tunneling dynamics of dipolar nets, when o


ξo , Vo (incoherent tunneling
relaxation regime). In (a), we show the short-time evolution of the distribution function M(ξ, t) =
P↑ (ξ, t) − −P↓ (ξ, t), where Pσ (ξ, t) is the normalized probability that a spin in a state |σ  = | ↑
or | ↓ finds itself in a bias field ξ at time t. Different colors show the distribution at different times.
At short times, a “tunneling hole” of width ξo appears, driven by inelastic tunneling transitions
involving nuclear spins. The dipolar interactions gradually modify the shape and width of the hole
at later times. This figure was produced by Monte Carlo simulations for a sample starting in a
strongly annealed state. (b) shows measurements of the function M(ξ = o , t = 0) on a strongly
annealed F e-8 crystal (from Ref. [30]), obtained by extracting the square root relaxation rate
−1
Γsqrt ≡ τQ ∼ (2o /Vo )ξo N (o ), where N (ξ ) is the “density of states” of spins in a bias energy ξ
(see text). If one lets M(ξ, t) relax for a time tW before examining it, the tunneling hole is revealed.
Closer examination (lower graph) shows that for small initial magnetization Min (i.e., strong
annealing), the hole has an intrinsic linewidth, revealed at short waiting times tW . This linewidth
ξo is caused by the nuclear spins (see text). (c) shows the tunneling matrix element extracted from
measurements of relaxation in a transverse field H⊥ o (from ref. [31]). These experiments found the
−1
oscillatory dependence of τQ , and thence |o |, on H⊥o (compare text, and Fig. 1c)

to relax, with a characteristic “square root” relaxation [27] ∼ (t/τQ )1/2 . One gets
−1
τQ (o ) ∼ (2o /Vo )ξo N(o ), where N(ξ ) is the “density of states” of spins in a bias
energy ξ , i.e., ξo N (H ) is the number of spins in the tunneling window, centered at
the external field bias energy o = gμB SHoz .
Many experiments, using ensembles of magnetic nanomolecules such as F e-8
and Mn-12 (which truncate to two-level systems at low T ), have now tested this
theory. Square root relaxation was found at short times [28, 29, 30]. Wernsdorfer et
−1
al. [30] found the time-evolving hole, of width ξo , by measuring τQ (ξ = o ) for
many different values of o (Fig. 3b). In strongly annealed samples (where M(ξ )
is a known Gaussian independent of sample shape), they also extracted 2o from
measurements of τQ , and showed how it oscillated in a transverse field (Fig. 3c),
and then confirmed this in independent “Landau-Zener” relaxation measurements.
We emphasize these oscillations are not evidence for coherent tunneling, quite the
5 Quantum Magnetism 271

contrary – the experiments observe incoherent relaxation rates! In a very striking


result [32], the nuclear isotopes were varied (substituting 2 H for 1 H , or 57 F e for
56 F e). This changed the hole width and the relaxation rate, giving independent

measurements of ξo which were consistent with the calculated value. This is fairly
direct evidence for the nuclear spin-mediated tunneling mechanism. The Leiden
group [33] has also done low-T NMR on Mn-12, seeing not only how nuclear
spins control the tunneling dynamics but also how the nuclear dynamics in turn is
controlled by the molecular tunneling dynamics.
Notice these are all results for short-time quantum relaxation. At longer times,
multi-spin correlations intervene, causing a breakdown of the square root – the
system moves into the quantum spin glass regime, of fundamental interest [35, 36].
Only quantum tunneling, simultaneously involving many spins, allows the system
to escape local potential minima. Experiments on the insulator LiH o0.44 Y0.56 F4 , in
which the lowest magnetic doublets (i.e., two-level systems) of J = 8 H o3+ ions
interact primarily via dipolar interactions, have looked at this tunneling relaxation
(Fig. 4a). Remarkably, much of the relaxation (here to a ferromagnetic ground state)
goes via collective dissipative tunneling of domain walls (see Fig. 4a); this purely
quantum effect has been definitively confirmed by observing its dependence on an
applied transverse field [37]. For these dense H o arrays, we can also reinterpret
the long-time relaxation as a quantum optimization process. One relaxes the system
toward the ground state not by reducing the temperature (as in thermal annealing
optimization protocols [38]), but by “quantum annealing” – exposing the system at
very low T to a large transverse field H⊥ o , allowing it to quantum relax, and then
reducing H⊥ o to zero, thence freezing the dynamics [39]. One then reads the final
state – which is the “solution” to the problem of energy optimization.
Finally, one can also explore the regime where ξo Vo , i.e., where dipolar
interspin interactions are unimportant, and the nuclear environment dominates
completely. Here, a “toothcomb” structure was expected in the quantum relaxation
rate, reflecting the level structure of nuclear spins [25]. This was recently found
(Fig. 4b) in experiments [40] on dilute concentrations of H o ions in LiH ox Y1−x F4
(there was also an interesting catch – residual inter-H o interactions can cause pairs
of spins to co-flip, giving a doubling of the teeth).
We see that the study of the quantum relaxation of a spin net reveals the essential
physics governing the incoherent spin dynamics. Now we can turn to the coherent
spin dynamics.

Large-Scale Coherence and Entanglement

The acid test of our understanding comes with large-scale quantum effects – where
decoherence must be rigorously suppressed. Of course the traditional view has been
that too many microstates are involved in any macrostate for this to be possible [41].
However, the modern picture is different.
Macroscopic tunneling: In pioneering work, predictions of macroscopic tun-
neling between different flux states in superconducting SQUID rings [20] were
272 G. Aeppli and P. Stamp

Fig. 4 (Continued)
5 Quantum Magnetism 273

c 70

60 0.11 K
0.11 K
χ'' (emu/mole Ho)
with pump
50
0.15 K

40

30

20

10

0
1 10 100 1000
f (Hz)
Fig. 4 Tunneling dynamics of the LiH ox Y1−x F4 system. In (a), we show typical behavior for
the rate of microscopic domain wall depinning in LiH o0.44 Y0.56 F4 , as a function of inverse
temperature (from Ref. [37]); the crossover from thermal activation to T -independent tunneling
−1
relaxation occurs when T ∼ 50 mK. (b) shows the relaxation rate τQ (H ) of the H o spins in a
−3
very dilute system (x = 2 × 10 ), from Ref. [40]. The main peaks in the “toothcomb” pattern,
each separated by the H o hyperfine energy, come from nuclear spin-mediated tunneling of single
H o ions between the lowest doublet states. The n = 8, 9 peaks shown in the inset come from co-
flip tunneling of pairs of H o ions, mediated by residual inter-H o interactions. (c) shows spectral
hole burning for x=0.045, from Ref. [57]. The absorptive part of the magnetic susceptibility is
measured as a function of frequency in the linear response regime using a probe amplitude of 0.04
Oe, giving a broad maximum centered at a frequency which depends strongly on temperature. The
same spectroscopy in the presence of a 0.2 Oe pump at 5 Hz shifts the spectrum, cuts off its tails,
and most important, inserts a sharp hole at 5 Hz.

quantitatively verified in the 1980s [42, 43]. In magnets, similar tunneling was
predicted for large domain walls pinned to defects [44], and also found in
experiments [45,46]. In experiments on large domain walls in mesoscopic Ni wires
(of thickness ∼20 − 80 nm), one sees a crossover to a T -independent escape rate
of the walls from a pinning center, in an applied field; the dependence of the rate
on field can be compared with theory [46]. This tunneling involved roughly 107
spins – not far short of the number of Cooper pairs involved in SQUID tunneling.
A revealing set of experiments [46] also looked at microwave absorption between
different levels – these represented the quantized dynamics of the wall center of
mass, trapped in the pinning potential well. A fairly detailed picture can be given of
these experiments [47, 48].
Such tunneling phenomena involve a collective degree of freedom (SQUID flux,
magnetic domain wall position) which does indeed involve a huge number of
274 G. Aeppli and P. Stamp

microstates. So how can it happen? One reason is that all electronic microstates
(Bogoliubov quasiparticles, magnons) are strongly gapped, by energies EG ∼
10 K. To such high-energy excitations, the collective coordinate tunneling, over a
timescale τo , seems very slow. The amplitude to excite them is then exponentially
small, ∼ O(e−EG τo ), by elementary time-dependent perturbation theory. Of course
there are also many very low-energy excitations (defects, paramagnetic impurities,
nuclear spins, etc. – the “satellite spins” discussed in section “The Importance
of Decoherence”), which can entangle with the collective tunneling coordinate.
However, they cause little dissipation, because their energy is so low – their direct
effect on “single-shot” tunneling is then rather weak.
Large-scale coherence: Coherent superpositions, on the other hand, require phase
coherence between many successive tunneling events [21,22]. Now, the low-energy
environmental microstates are indeed very dangerous [19]. So is macroscopic state
superposition feasible?
In superconductors, the answer to this question came a considerable time ago,
including early experimental evidence for macroscopic flux state superpositions
[49, 50, 51]. The decoherence times τφ for superconducting qubits have undergone
a spectacular rise since the year 2000 so that today, they are viewed as leading
candidates for the fundamental building blocks of quantum computers.
Analogous macroscopic superpositions in magnets - e.g., of “giant qubit”
superpositions of two different magnetization states in a large magnetic particle –
have not yet been confirmed. Some years ago, absorption experiments in very large
ferritin molecules (with Neél vector ∼ 23, 000 μB ) showed sharp resonances,
attributed to coherent tunneling of individual ferritin molecules [53]. However, these
results were hard to understand theoretically, and no other group has confirmed
them; and experiments on much smaller molecules like Mn-12 or F e-8 have
never seen coherence. The basic difficulty is that low-energy decoherence from the
hyperfine coupling to nuclear spins is expected to be large (in ferritin, the hyperfine
coupling to a single 57 F e nucleus is much larger than the tunnel splitting!), and at
higher energies, phonon contributions are not negligible [25, 19].
However, experiments may have simply been looking in the wrong place. The
“window of opportunity” between nuclear spin and phonon energy scales is actually
very wide; by applying strong transverse fields, one can increase o to values
much higher than hyperfine couplings (compare Fig. 1c), but still much lower than
most phonon energies, and optimize the decoherence rate. By combining isotopic
purification with choice of material, one can also remove almost all nuclear spins.
Experimentalists like to define a decoherence Q-factor Qφ = o τφ , which tells us
how many coherent oscillations the system can show before decoherence sets in.
Elementary theory [25, 52] then indicates that for an spin S in an insulator, we have
an optimal Qφ ∼ θD 2 /SK E , where K is the anisotropy energy per electronic spin
o o o
of the magnet, θD the Debye energy, and Eo is again the spreading of the nuclear
multiplet. For example, if θD = 300 K, Ko = 1 K, and S = 106 , then by reducing
Eo to
0.1 K, we should get “mesoscopic” coherent dynamics (i.e., Qφ > 1).
Large-scale entanglement: We next turn to multi-qubit entangled states, but now
involving microscopic spins. It is estimated that quantum information processing
5 Quantum Magnetism 275

Fig. 5 Design for a nuclear


spin-based quantum
computer, from Kane [54].
Two cells in a
one-dimensional array,
containing 31P donors and
electrons in a Si host, are
separated by a barrier from
metal gates on the surface.
The “A gates” control the
resonance frequency of the
nuclear spin qubits, while the
“J gates” control the
electron-mediated coupling
between adjacent nuclear
spins

(QUIP) will be possible, using error correction [14,16], if the single qubit coherence
Q-factor Qφ > 104 − 105 . This should easily be possible with microscopic spins –
note from above that the optimal Qφ ∼ O(1/S). Thus, theory clearly indicates
that QUIP is feasible with microscopic magnetic qubits, provided electronic deco-
herence is absent (e.g., magnetic ions in insulators or semiconductors, or perhaps
insulating molecular crystals). Many proposals have appeared along these lines – as
an example, consider that due to Kane [54] in Fig. 5, using networks of nuclear spins
in semiconductors to do the computation. Reasonable estimates of decoherence rates
[54] give very small numbers here – problems should only arise, as before, from
very low-energy excitations (e.g., 1/f noise from charge defects). Again, applying
strong fields should help [55], and measurements of the decoherence rates will be
crucial, in this and other designs [56].
Experiments over the last years have given very long spin relaxation times for the
spins associated with isolated impurities and quantum dots in silicon. While this is
very interesting for quantum information science, the finding of sharp resonances in
a strongly interacting many-body system using spectral hole-burning [57, 64, 69]
in LiH o0.045 Y0.995 F4 (see Fig. 4c) is important for the science of disordered
magnetism. The decoherence was remarkably small, in spite of the long-range inter-
H o dipole interactions. The data were explained as a collective effect, involving
tunneling of large clusters of H o spins. These results are both surprising and
exciting – they indicate that we may be close to manipulating entangled mesoscopic
spin states.
To do fully fledged quantum computations will require controlling individual
spins or groups of spins, i.e., control of the parameters i , i , and Vij . Control of i
and i can obviously be done by varying transverse and longitudinal external fields –
in fact, all quantum logic operations can be implemented by varying only one of
these three parameters, and one can also use timed pulses in creative ways (which
also help with decoherence [58,59]), so control of Vij is not crucial. One possibility
276 G. Aeppli and P. Stamp

is to use engineered heterostructures to control the local fields [54]; another would
be to use magnetic STM tips, although the practicality of this for anything other than
demonstration experiments involving a very small number of qubits is questionable.
A more difficult problem will be to measure the quantum state of the spin qubits,
without affecting their operation. One can imagine many possibilities – for example,
bringing in superconducting or magnetic devices, whose tunneling into the qubit
depends on its polarization [60], or using optical methods. Over the last two decades,
great progress has been made on both adiabatic and gated quantum computation.
For solid-state implementations, the leading contenders have been superconducting
qubits, whose decoherence times have dramatically improved, and for which very
complex circuits can be constructed. It is beyond the scope of this chapter on
magnetism to review these developments on magnetism, except to mention that
the quantum annealer manufactured by D-wave systems [62] is built to simulate
the transverse field Ising model, and can therefore be viewed as a programmable
version of LiH o1−x Yx F4 . Equally relevant here and for the future of magnetism are
experiments showing control of the magnetic interactions between the very simple
S=1/2 spins associated with either donors [34] or quantum dots [67] in silicon.

Future Directions and Open Problems

When many of the authors of this volume began graduate work in the 1980s, the
idea of large-scale quantum phenomena in magnetic systems was hardly a topic for
discussion – for ∼70 years, quantum mechanics was only used in magnetism to
discuss atomic and nuclear spins and the microscopic interactions (exchange, spin-
orbit, etc.) operating on them. Now we are discussing and even seeing quantum
phenomena at much larger scales, where magnetic variables were previously treated
as classical – it is in this sense that the “quantum” is being put back into magnetism.
We are on the threshold of a very different era – in which coherent spin states, having
no classical analogue, may come to play a role as important as the macroscopic wave
function in superconductors.
As always, it is difficult to make predictions about a fast-moving field. The
preparation and readout of coherent multi-spin states may well require techniques
from spintronics [61], implemented on submicron scales. The key challenges
here will be (i) to marry spintronics with the science of collective quantum spin
states, under progressively less extreme experimental conditions of magnetic field
and temperature, and (ii) to understand how to suppress electronic decoherence
in conducting magnets. This latter is a hard problem because spin current is not
a conserved quantity, which has made it difficult to find a rigorous theory of
spin dynamics in conducting magnets, although moving to antiferromagnets where
mesoscopic quantum tunneling of domain walls has been identified (in the common
metal chromium) [68] may be a promising route. The development of new materials
having the correct mix of optical, electronic, and magnetic properties will be crucial,
and theory will need to be developed to model candidate materials and devices.
It is sobering that even for an insulator as simple as LiH ox Y1−x F4 , many key
5 Quantum Magnetism 277

discoveries – e.g., the coherent hole burning for x=0.045 – were unexpected, and
dictated by such factors as the availability of samples with particular compositions
at sale prices.
We also note that there are many interesting spin systems apart from electronic
magnets. Magnetic superfluids like 3 H e (see, e.g., [63]), or spin-1 Bose-Einstein
condensates (BECs) of alkali atoms [65, 66], and quantum Hall spin condensates,
offer examples where a spin coherent wave function coupled to another order
parameter can display very rich dynamics, and where large-scale quantum effects
are worth exploring.
Finally, we emphasize a crucial change of perspective in the field. Condensed
matter physicists are accustomed to dealing with only one- and two-spin correlation
functions, whereas quantum information theory requires manipulation and measure-
ment of multi-qubit correlations. We still have a great deal to learn about these, and
the next decades promise to be a very exciting one for quantum magnetism.

References
1. London. F.: The λ-phenomenon of liquid H e and the Bose-Einstein degeneracy. Nature 141,
643 (1938)
2. Feynman, R.P.: Spacetime approach to non-relativistic quantum mechanics. Rev. Mod. Phys.
20, 367 (1948)
3. Feynman, R.P.: The Feynman Lectures on Physics, vol. 3. Addison-Wesley (1965)
4. Shapere. A., Wilczek. F.: Geometric Phases in Physics. World Scientific (1989)
5. Klauder, J.R.: Path integrals and stationary phase approximations. Phys. Rev. D19, 2349 (1979)
6. Berry, M.V.: Quantal Phase factors accompanying adiabatic changes. Proc. Roy. Soc. A392, 45
(1984)
7. Auerbach. A.: Interacting Electrons and Quantum Magnetism. Springer (1994)
8. Aharonov. Y, Bohm. D.: Significance of electromagnetic potentials. Phys. Rev. 115, 485 (1959)
9. Haldane, F.D.M.: Non-linear field thery of large-spin Heisenberg antiferromagnets: semiclas-
sically quantized solitons of the 1-d easy-axis Néel state. Phys. Rev. Lett. 50, 1153 (1983)
10. Bogachek, E.N., Krive, I.V.: Quantum oscillations in small magnetic particles. Phys. Rev. B46,
14559 (1992)
11. Garg. A.: Topologically Quenched Tunnel Splitting in Spin Systems without Kramers’
Degeneracy. Europhys. Lett. 22, 205 (1993)
12. Tupitsyn, I.S., Barbara, B.: Quantum tunneling of Magnetisation in molecular complexes with
large spins: effect of the Environment. In: Miller, J.S., Drillon, M. (eds.) Magnetoscience- from
Molecules to Materials, vol. 3, pp. 109–168. Wiley (2001)
13. Kramers. H.A.: Théorie générale de la rotation paramagnétique dans les cristaux. Proc. Acad.
Sci. Amst. 33, 959 (1930)
14. Bennett, C.H., DiVincenzo, D.: Quantum information and computation. Nature 404, 247
(2000)
15. Lo, H.-K., Popescu, S., Spiller, T.: Introduction to Quantum Computation and Information.
World Scientific (1998)
16. Nielsen, M.A., Chuang, I.L.: Quantum Computation and Quantum Information. Cambridge
University Press (2000)
17. Mohanty, P., Jariwala, E.M.Q., Webb, R.A.: Intrinsic decoherence in mesocopic systems. Phys.
Rev. Lett. 78, 3366 (1997)
18. Aleiner, I.L., Altshuler, B.L., Gershenson, M.E.: Interaction effects and phase relaxation in
disordered systems. Waves Random Media 9, 201 (1999)
278 G. Aeppli and P. Stamp

19. Prokof’ev, N.V., Stamp, P.C.E.: Theory of the spin bath. Rep. Prog. 63, 669 (2000)
20. Caldeira, A.O., Leggett, A.J.: Quantum tunneling in a dissipative system. Ann. Phys. 149, 374
(1984)
21. Leggett, A.J., Chakravarty. S, Dorsey, A.T., Fisher, M.P.A., Garg, A., Zwerger, W.: Dynamics
of the dissipative 2-state system. Rev. Mod. Phys. 59, 1 (1987)
22. Weiss, U.: Quantum Dissipative Systems. World Scientific (1999)
23. Prokof’ev, N.V., Stamp, P.C.E.: Giant spins and topological decoherence: a Hamiltonian
approach. J. Phys. CM 5, L663 (1993)
24. Anderson, P.W.: Infrared catastrophe in fermi gases with local scattering potentials. Phys. Rev.
Lett. 18, 1049 (1967).
25. Prokof’ev, N.V., Stamp, P.C.E.: Quantum relaxation of magnetisation in magnetic Particles. J.
Low Temp. Phys. 104, 143 (1996)
26. Dubé, M., Stamp, P.C.E.: Mechanisms of decoherence at low Temperatures. Chem. Phys. 268,
257 (2001)
27. Prokof’ev, N.V., Stamp, P.C.E.: Low-T relaxation in a system of Magnetic molecules. Phys.
Rev. Lett. 80, 5794 (1998)
28. Ohm, T., et al.: Local field dynamics in a resonant quantum tunneling system of magnetic
molecules. Europhys. J. B6, 595 (1998)
29. Thomas, L., et al.: Non-exponential scaling of the magnetisation relaxation in Mn12 acetate.
Phys. Rev. Lett. 83, 2398 (1999)
30. Wernsdorfer, W., Ohm, T., Sangregorio, C., Sessoli, R., Mailly, D., Paulsen, C.: Observation of
the distribution of molecular spin states by resonant quantum tunneling of the magnetisation.
Phys. Rev. Lett. 82, 3903 (1999)
31. Wernsdorfer, W., Sessoli, R.: (1999) Quantum phase interference and parity effects in Magnetic
molecular clusters. Science 284, 133
32. Wernsdorfer, W., Caneschi, A., Sessoli, R., Gatteschi, D., Cornia, A., Paulsen, C.: Effect of
nuclear spins on the quantum relaxation of the magnetisation for the molecular magnet F e-8.
Phys. Rev. Lett. 84, 2965 (2000)
33. Morello, A., et al.: Quantum tunneling of magnetisation in Mn12-ac studied by Mn-55 NMR.
Polyhedron 22, 1745–1749 (2003)
34. Veldhorst, M., Yang, C.H., Hwang, J.C.C., Huang, W., Dehollain, J.P., Muhonen, J.T.,
Simmons, S., Laucht, A., Hudson, F.E., Itoh, K.M., Morello, A., Dzurak, A.S.: A two-qubit
logic gate in silicon. Nature 526, 410(2015)
35. Sachdev, S.: Quantum Phase Transitions. Ch. 16. Cambridge University Press (1999)
36. Thill, M.J., Huse, D.A.: Equilibrium behaviour of quantum Ising spin glass. Physica 214A, 321
(1995)
37. Brooke, J., Rosenbaum, T.F., Aeppli, G.: Tunable quantum tunneling of magnetic domain
walls. Nature 413, 610 (2001)
38. Kirkpatrick, S., Gelatt, C.D., Vecchi, M.P.: Optimisation by simulated annealing. Science 220,
671 (1983)
39. Brooke, J., Bitko, D., Rosenbaum, T.F., Aeppli, G.: Quantum annealing of a disordered magnet.
Science 284, 779 (1999)
40. Giraud, R., Wernsdorfer, W., Tkachuk, A.M., Mailly, D., Barbara, B.: Nuclear spin driven
quantum relaxation in LiY0.998 H o0.002 F4 . Phys. Rev. Lett. 87, 057203 (2001)
41. van Kampen, N.G.: Ten Theorems about Quantum Mechanical Measurements. Physica A153,
97 (1988)
42. Washburn, S., Webb, R.A., Voss, R.F., Faris, S.M.: Effects of dissipation and temperature on
macroscopic quantum tunneling. Phys. Rev. Lett. 54, 2712 (1985)
43. Clarke, J., Cleland, A.N., Devoret, M., Esteve, D., Martinis, J.M.: Quantum mechanics of
a macroscopic variable: the phase difference of a Josephson junction. Science 239, 992
(1998)
44. Stamp, P.C.E.: Quantum dynamics and tunneling of domain walls in ferromagnetic insulators.
Phys. Rev. Lett. 66, 2802 (1991)
45. Paulsen, C., et al.: Macroscopic Quantum Tunnelling Effects of Bloch Walls in Small
Ferromagnetic Particles. Europhys. Lett. 19, 643 (1992)
5 Quantum Magnetism 279

46. Hong, K., Giordano, N.: Effect of microwaves on domain wall motion in this Ni wires.
Europhys. Lett. 36, 147 (1996)
47. Tatara, G., Fukuyama, H.: Macroscopic quantum tunneling of a domain wall in a ferromagnetic
metal. Phys. Rev. Lett. 72, 772 (1994)
48. Dubé, M., Stamp, P.C.E.: Effect of Phonons and Nuclear Spins on the tunneling of a Domain
wall. J. Low Temp. Phys. 110, 779 (1998)
49. Van der Wal, C.H., ter Haar, A.J.C., Wilhelm, F.K., Shouten, R.N.: Harmans, C.J.P.M.,
Orlando, T.P., Lloyd, S., Mooij, J.E.: Quantum Superposition of Macroscopic Persistent-
Current states. Science 290, 773 (2000)
50. Vion, D., Aasime, A., Cottet, A., Joyez, P., Pothier, H., Urbina, C., Esteve, D., Devoret, M.:
Manipulating the Quantum State of an Electrical Circuit. Science 296, 886 (2002)
51. Chiorescu, I., Nakamura, Y., Harmans, C.J.P.M., Mooij, J.E.: Coherent quantum dynamics of
a superconducting flux qubit. Science 299, 1869 (2003)
52. Stamp, P.C.E., Tupitsyn, I.S.: Coherence window in the dynamics of quantum nanomagnets.
Phys. Rev. B69, 014401 (2004)
53. Gider, S., et al.: Classical and quantum magnetic phenomena in natural and artificial ferritin
proteins. Science 268, 77 (1995)
54. Kane, B.L.: A Si-based nuclear spin quantum computer. Nature 393, 133 (1998)
55. Privman, V., Vagner, I.D., Kventsel, G.: Quantum computation in quantum-Hall systems Phys.
Lett. A236, 141 (1998)
56. Barrett, S.D., Milburn, G.J.: Measuring the decoherence rate in a semiconductor charge qubit.
Phys. Rev. B68, 155307 (2003)
57. Ghosh, S., Parthasarathy, R., Rosenbaum, T.F., Aeppli, G.: Coherent spin oscillations in a
disordered magnet. Science 296, 2195 (2002)
58. Viola, L., Lloyd, S.: Dynamical supression of decoherence in 2-state quantum systems. Phys.
Rev. A58, 2733 (1998)
59. Wu, L.A., Lidar, D.A.: Creating decoherence free subspaces using strong and fast pulses. Phys.
Rev. Lett. 88, 207902 (2002)
60. Manoharan, H.C., Lutz, C.P., Eigler, D.M.: Quantum Mirages formed by coherent projection
of electronic structure. Nature 403, 512 (2000)
61. Wolf, S.A., Awschalom, D.D., Buhrman, R.A., Daughton, J.M., von Molnar, S., Roukes, M.L.,
Chtchelkanova, A.Y., Treger, D.M.: Spintronics: a spin-based electronics vision for the future.
Science 294, 1488 (2001)
62. Johnson, M.W., Amin, M.H.S, Gildert, S., Lanting, T., Hamze, F., Dickson, N., Harris, R.,
Berkley, A.J., Johansson, J., Bunyk, P., Chapple, E.M., Enderud, C., Hilton, J.P., Karimi, K.,
Ladizinsky, E., Ladizinsky, N., Oh, T., Perminov, I., Rich, C., Thom, M.C., Tolkacheva, E.,
Truncik, C.J.S., Uchaikin, S., Wang, J., Wilson, B., Rose, G.: Quantum annealing with
manufactured spins. Nature 473 194 (2011)
63. Vollhardt, D., Wölfle, P.: The superfluid phases of Helium 3. Taylor & Francis (1990)
64. Schmidt, M.A., Silevitch, D.M., Aeppli, G., Rosenbaum, T.F.: Using thermal boundary
conditions to engineer the quantum state of a bulk ferromagnet. PNAS 111, 3689 (2014)
65. Stamper-Kurn, D.M., Miesner, H.J., Chikkatur, A.P., Inouye, S., Stenger, J., Ketterle, W.:
Quantum tunneling across spin domains in a BEC. Phys.Rev. Lett. 83, 661 (1999)
66. Ketterle, W.: Spinor condenstates and light scattering from Bose-Einstein condensates. Proc.
Les Houches summer school LXXII (1999)
67. Watson, T.F., Philips, S.G.J., Kawakami, E., Ward, D.R., Scarlino, P., Veldhorst, M., Savage,
D.E., Lagally, M.G., Mark Friesen, Coppersmith, S.N., Eriksson, M.A., Vandersypen, L.M.K.:
A programmable two-qubit quantum processor in silicon. Nature 555, 263 (2018)
68. Shpyrko, O., Isaacs, E., Logan, J., Yejun Feng, Aeppli. G, Jaramillo, R., Kim, H.C.,
Rosenbaum, T.F., Zschack, P., Sprung, M., Narayanan, S., Sandy, A.R.: Direct measurement
of antiferromagnetic domain fluctuations. Nature 447, 68–71 (2007). https://doi.org/10.1038/
nature05776
69. Silevitch, D.M., Tang, C., Aeppli, G., Rosenbaum, T.F.: Tuning high-Q nonlinear dynamics in a
disordered quantum magnet. Nat. Commun. 10, 4001 (2019). https://doi.org/10.1038/s41467-
019-11985-1
280 G. Aeppli and P. Stamp

Gabriel Aeppli is professor at ETHZ and EPFL, and head of


Photon Science at PSI. After working at NEC, AT&T, IBM and
MIT on problems from liquid crystals to magnetic data storage,
he co-founded the London Centre for Nanotechnology and Bio-
Nano Consulting. His focus is on implications and development
of photon science and nanotechnology for information processing
and health care.

Philip Stamp received his PhD from the Univ of Sussex. After
postdoctoral work in Massachusetts, Grenoble, and Santa Bar-
bara, he held positions in the Univ of British Columbia, and as
a Spinoza Professor in Utrecht. He is currently director of the
Pacific Institute of Theoretical Physics in Vancouver. He works
on theoretical quantum gravity and condensed matter theory.
Spin Waves
6
Sergej O. Demokritov and Andrei N. Slavin

Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 282
Spin Waves in 3D and 2D Systems: Theory and Experiment . . . . . . . . . . . . . . . . . . . . . . . . . . 284
Theory of Spin Waves in 3D and 2D Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 284
Brillouin Light Scattering a Powerful Tool for Investigation of Spin Waves . . . . . . . . . . . . 288
Spin Waves in 1D Magnetic Elements: Standing and Propagating Waves . . . . . . . . . . . . . . . . 291
BLS in Laterally Confined Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 291
Lateral Quantization of Spin Waves in Magnetic Stripes . . . . . . . . . . . . . . . . . . . . . . . . . . . 291
Spin Wave Wells and Edge Modes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 296
Implementation of Micro-Focus BLS for Laterally Patterned Magnetic Systems . . . . . . . . 300
Propagating Waves in 1D Magnetic Structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 302
Control and Conversion of the Propagating Waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 307
Inductive Excitation of Spin Waves in 1D Waveguides . . . . . . . . . . . . . . . . . . . . . . . . . . . . 311
Spin-Torque Transfer Effect and Spin Waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 315
Spin Waves in 0D . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 320
Spin-Torque Nano-Oscillator (STNO) and Emitted Spin Waves . . . . . . . . . . . . . . . . . . . . . 321
Spin-Hall Nano-Oscillator (SHNO) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 324
Nature of Spin Wave Modes Excited in 0D Magnetic Nanocontacts . . . . . . . . . . . . . . . . . . 329
Coupling of a STNO and 1D Spin-Wave Waveguide to Each Other . . . . . . . . . . . . . . . . . . 336
Conclusion and Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 339
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 341

S. O. Demokritov ()
Institute for Applied Physics and Center for Nanotechnology, University of Muenster, Muenster,
Germany
e-mail: demokrit@uni-muenster.de
A. N. Slavin
Department of Physics, Oakland University, Rochester, MI, USA
e-mail: slavin@oakland.edu

© Springer Nature Switzerland AG 2021 281


J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_6
282 S. O. Demokritov and A. N. Slavin

Abstract

Spin waves are the dynamic eigen-excitations of a magnetic system. They


provide the basis for the description of spatial and temporal evolution of the
magnetization distribution of a magnetic object. The unique features of spin
waves such as the possibility to carry spin information over relatively long
distances, the possibility to achieve submicrometer wavelength at microwave
frequencies, and controllability by electronic signal via magnetic fields make
these waves uniquely suited for implementation of novel integrated electronic
devices characterized by high speed, low power consumption, and extended
functionalities. The history of spin waves clearly shows a progressively increas-
ing interest for the spin waves in magnetic samples of reduced dimensionality.
Since 1950s the focus of the researchers has moved from 3D to 2D objects –
thin films and magnetic multilayers, resulting in the discovery of the surface
Damon-Eshbach spin-wave mode. Later in 1990s 1D stripes became the most
actively studied magnetic systems, bringing about the discovery of lateral
quantization and edge modes. Finally, theoretical prediction of the spin-torque
effect and development of novel techniques for nanofabrication allowed for the
investigation of magnetic 0D objects such as spin-torque nano-oscillator. In this
chapter we follow this historical trend and describe the recent development of
spin-wave studies.

Introduction

Felix Bloch introduced the concept of spin waves (SPW), as the lowest-energy mag-
netic states above the ground state of a magnetic medium [1]. Bloch theoretically
considered quantum states of magnetic systems with spins slightly deviating from
their equilibrium orientations, and found that these disturbances were dynamic: they
propagate as waves through the medium. It is interesting to note that the concept
of dynamic spin waves was introduced to explain experimental data obtained
in static measurements. From spin-wave theory Bloch was able to predict that
the magnetization of a three-dimensional (3D) ferromagnet at low temperatures
should deviate from its zero-temperature value with a T3/2 dependence (the famous
Bloch law), instead of the exponential dependence given by the mean field theory.
Albeit the Bloch theory has restricted itself by assuming the dominating exchange
interaction, now we know that the relativistic magnetic dipole interaction plays
a decisive role in the properties of spin waves having the wavelengths that are
much smaller than the interatomic distance in the magnetic medium. Moreover, phe-
nomenologically, spin waves in a wide interval of wavevectors (30 < k < 106 cm−1 )
that is most important for the practical applications are, on one hand, almost entirely
determined by the magnetic dipole-dipole interaction, and, on the other hand, can
be correctly described when the retardation effects are neglected. Such spin waves
are usually called dipolar magnetostatic waves or magnetostatic modes. Due to the
anisotropic properties of the magnetic dipole interaction, the frequency of a spin
6 Spin Waves 283

wave depends on the orientation of its wavevector relative to the orientation of


the static magnetization. For higher values of the wavevectors, when the exchange
interaction cannot be neglected, one speaks about dipole-exchange spin waves.
Spin waves are the dynamic eigen-excitations of a magnetic system. They
provide the basis for the description of spatial and temporal evolution of the
magnetization distribution of a magnetic object under the general assumption that
locally the length of the magnetization vector is constant. This is correct, if, first,
the temperature is far below the Curie temperature of the medium, as is assumed
throughout this chapter, and, second, if no topological anomalies such as vortices or
domain walls are present. The latter is fulfilled for samples in a single-domain state,
i.e., magnetized to saturation by an external bias magnetic field.
The unique features of spin waves such as the possibility to carry spin infor-
mation over relatively long distances, the possibility to achieve submicrometer
wavelength at microwave frequencies, and controllability by electronic signal via
magnetic fields make these waves uniquely suited for implementation of novel
integrated electronic devices characterized by high speed, low power consumption,
and extended functionalities. The utilization of spin waves for integrated electronic
applications is addressed within the emerging field of magnonics [2–5]. Although
the application of spin waves for microwave signal processing has been intensively
explored since many decades, recent advances in spintronics and nanomagnetism, as
well as the development of novel techniques for nanofabrication and measurements
of high-frequency magnetization dynamics created essentially new possibilities for
magnonics and brought it onto a new development stage. Of particular importance
here is the recent discovery of the spin-transfer torque (STT) [6–8] and the spin-Hall
effect (SHE) [9–11], both of which have already been demonstrated to enable novel
device geometries and functionalities [12–16].
The history of spin waves clearly shows a progressively increasing interest
to the spin waves in magnetic samples of reduced dimensionality. Although the
original theory of Bloch was developed for 3D magnets, since 1950s the focus
of the researchers has moved to two-dimensional (2D) objects – thin films and
magnetic multilayers, resulting in the discovery of quantized standing spin-wave
resonances [17, 18], surface Damon-Eshbach spin-wave mode [19, 20], and the
interlayer coupling [21]. Later in 1990s quasi-one-dimensional (1D) stripes became
the most actively studied magnetic systems, bringing about the discovery of
lateral quantization [22] and edge modes [23, 24]. Finally, as mentioned above,
theoretical prediction of the spin-torque effect (STE) [6, 7] and development of
novel techniques for nanofabrication allowed for investigation of magnetic zero-
dimensional (0D) objects such as spin-torque nano-oscillator (STNO) [25–31]. In
this chapter we follow this historical trend. In the Sect. “Spin Waves in 3D and 2D
Systems: Theory and Experiment” we describe the theory of spin waves in 3D and
2D magnetic systems, as well as the main experimental techniques for spin-wave
studies in such systems. In Sect. “Spin Waves in 1D Magnetic Elements: Standing
and Propagating Waves” we focus on spin waves in quasi-1D systems: laterally
quantized and localized spin-wave edge modes in magnetic stripes. We also discuss
the propagating wave modes in magnetic waveguides. The experimental data are
284 S. O. Demokritov and A. N. Slavin

complemented by a short theoretical description of the spin-torque effect and its


role in the damping compensation for propagating spin waves. In Sect. “Spin Waves
in 0D” we describe the spin-wave dynamics of 0D systems on the example of an in-
plane magnetized STNO structure supporting a self-localized solitonic spin-wave
“bullet” mode. Conclusions are given in Sect. “Conclusion and Outlook”.

Spin Waves in 3D and 2D Systems: Theory and Experiment

This section comprises two subsections. Subsection “Theory of Spin Waves in 3D


and 2D Systems” is devoted to the general theory of spin waves in both 3D and
2D systems. It also demonstrates the modification of the spin-wave properties due
to the dimensionality reduction. Subsection “Brillouin Light Scattering a Powerful
Tool for Investigation of Spin Waves” describes the Brillouin light scattering – the
main experimental techniques for the investigation of spin waves. Since spin waves
in 3D and 2D systems have been intensively studied in the past, the sections devoted
to their description are short, and the main attention is paid to the spin waves in the
quasi-1D and 0D geometries.

Theory of Spin Waves in 3D and 2D Systems

The dynamics of the magnetization vector is described by the Landau-Lifshitz


torque equation [32]:
1 dM
− = M × H eff (1)
γ dt

where M = MS + m(R, t) is the total magnetization, MS and m(R , t) are the vectors
of the saturation and the variable magnetization correspondingly, γ is the modulus
of the gyromagnetic ratio for the electron spin (γ/2π = 2.8 MHz/Oe), and

δW
H eff = − (1a)
δM
is the effective magnetic field calculated as a variational derivative of the energy
function W, where all the relevant interactions in the magnetic substance have
been taken into account (see, e.g., [33–35]). For the case of an unbounded 3D
ferromagnetic medium the variable magnetization m(R, t) depends on time t and
on the three-dimensional radius-vector R. In the spin-wave analysis it is usually
assumed that the variable magnetization m(R, t) is small compared to the saturation
magnetization MS , i.e., the angle of magnetization precession is small. In this case
the variable magnetization can be expanded in a series of plane spin waves (having
a 3D wavevector q):

m (R, t) = mq exp (iqR) . (2)
q
6 Spin Waves 285

The spectrum of dipole-exchange spin waves is in an unbounded ferromagnetic


medium which is given by the Herrings-Kittel formula [36].
  1/2
2A 2 2A 2
ω = 2πf = γ H+ q H+ q + 4π Ms sin2 θq (3)
Ms Ms

where A is the exchange stiffness constant, H is the applied magnetic field, and θ q
is the angle between the directions of the wavevector q and the static magnetization
MS with sin2 θ q being the matrix element of the dipole-dipole (magnetostatic)
interaction. Analyzing Eq. (3), one concludes that if the exchange can be neglected
q2 < < HMS /2A and q 2 << 2π MS2 /A, the spin-wave frequency is independent of q.
It depends solely on θ q demonstrating that the nonexchange (magnetostatic) spin-
wave spectrum is anisotropic and nondispersive. In contrast, for q2 > > HMS /2A
and q 2 >> 2π MS2 /A the spectrum of purely exchange spin waves is isotropic since
the spin wave frequency solely depends on q = |q|.
The transition from 3D to 2D can be made if one considers a magnetic film
with a finite thickness d. In the following, we assume a Cartesian coordinate system
oriented in such a way that the film normal is along the x-axis, and axes y and z are
in the film plane, with the external field H and the static magnetization MS being
aligned along the z-axis. Correspondingly, because the translational invariance along
the direction normal to the film surfaces (axis x) is broken, the three-dimensional
spin wave wavevector is represented as a sum of a two-dimensional continuous in-
plane wavevector q and quantized wavevector κ p ex (p = 0,1,2 . . . ) along the film
thickness: q = q + κ p ex , while the three-dimensional radius vector is represented
as R = R + x ex . Then, the distribution of the variable magnetization along the
film thickness (axis x) can be represented as a Fourier series in a complete set of
orthogonal functions mp (x) [37]:
  
m (R, t) = mp (x) exp iq  R  . (2a)
q  ,p

These functions mp (x) are chosen in such a way that they satisfy the exchange
differential operator of the second order and the exchange boundary conditions at
the film boundaries [38]:

∂m
+ D m|x=±d/2 = 0, (4)
∂x
x=±d/2

where D is the so-called “pinning” parameter determined by the ratio of the


effective surface anisotropy ks and the exchange stiffness constant A: D = ks /A.
The modes mp (x) can be interpreted as the modes of the spin-wave resonance
[17] in a particular geometry, and are sometimes called perpendicular standing
spin waves (PSSW). The discrete transverse wavenumbers κ p for these modes
are determined from the eigenvalue problem for the exchange differential operator
286 S. O. Demokritov and A. N. Slavin

with the boundary conditions Eq. (4). Note that in an finite-in-plane nonellipsoidal
magnetic film samples, the “pinning” of the dynamic magnetization at the lateral
edges of the sample could be determined by the local inhomogeneity of internal
dynamic magnetic field, and a different set of the “in-plane” eigenfunctions for the
expansion of the in-plane components of the variable magnetization can be obtained
in that case [39, 40].
Assuming that the thickness spin-wave modes mp (x) do not hybridize, it is
possible to obtain an approximate “diagonal” dispersion equation for the spin-wave
modes in a magnetic film of a finite thickness that is similar to the classical Kittel
equation (3) [37]:
 
2A
2
ωp = 2πfp = γ H + q + κp2
Ms 
  (5)
2A
2   1/2
H+ q + κp + 4π Ms Fpp κp , q , d
2
,
Ms 

where Fpp (κ p , q , d) is the matrix element of the dipole-dipole interaction in a film


defined by Eq. (46) in [37].
In the case of “unpinned” spins at the film surfaces

∂m
= 0, (6)
∂x
x=±d/2

it is possible to obtain a simple explicit expression for the transverse spin-wave


wavenumber κ p = pπ /d, p = 0, 1, 2, . . . , and the expression for the matrix
element Fpp (κ p , q , d) for an arbitrary angle between q|| and MS can be written
in the form:
    
4π MS qy2 q2
Fpp = 1 + Ppp (q) 1 − Ppp (q) − Ppp (q) z2 ,
H + (2A/MS ) q 2 q2 q
(7)

where

pπ 2
pπ 2
q 2 = qy2 + qz2 + = q2 + , (8)
d d

and the function Ppp (q , p) is defined in [37].


We present here only the expression for this function for the lowest (quasi-
uniform) thickness mode (p = 0) [37]:
 
  1 − exp −q d
P00 = P00 q d = 1 + . (9)
q d
6 Spin Waves 287

If the lowest thickness spin wave mode (p = 0) is propagating perpendicular to


the bias magnetic field, q = (qy , 0), its dispersion equation obtained from (5) using
(6), (7), and (8) in the nonexchange limit (A = 0) has the form:
   
ω0 qy d = 2πf0 qy d
    1/2 (10)
= γ H (H + 4π MS ) + (4π MS )2 P00 qy d 1 − P00 qy d ,

which for qy d < < 1 is similar to the dispersion equation obtained by Damon and
Eshbach for the dipolar surface mode [19]:
   
ωDE qy d = 2πfDE qy d
   1/2 (11)
= γ H (H + 4π MS ) + (2π MS )2 1 − exp −2qy d .

Thus, the spectrum of spin wave modes propagating perpendicular to the


direction of the bias magnetic field in an in-plane magnetized magnetic film obtained
from (5), (7), and (8) contains the lowest dipole-dominated mode (p = 0) with a
quasi-uniform thickness profile, which is analogous to the spin waves in 3D bulk
samples, and higher exchange-dominated spin-wave modes (p > 0), whose thickness
profiles are approximately described by the PSSWs. The higher spin-wave modes
are created because of the broken translational invariance along the x-direction. The
frequencies of these modes (p > 0) in the long wave limit qy d  1 can be calculated
from the following expression:
 
pπ 2 
2A
ωp = 2πfp = γ H + qy +
2
Ms d
  2 1/2 (12)
2A
pπ 2  
4π M /H
s
H+ qy2 + + 4π Ms + H qy2 ,
Ms d pπ/d

which is obtained from Eq. (5) using the expressions for the dipole-dipole matrix
elements Fpp (qy d) calculated in [37].
Figure 1 illustrates the typology of the lowest quasi-uniform (p = 0) dipole-
dominated spin-wave modes in the quasi-2D case of a magnetic film for different
mutual orientations between the in-plane wavevector, q , and the static magneti-
zation, MS . Three different geometries are shown. If MS is in the film plane, and
if q is perpendicular to MS , the surface or Damon-Eshbach (DE) mode exists (see
Eqs. (10) and (11)). If q and MS are collinear in the film plane, a mode with a
negative dispersion, or, so-called, the backward volume magnetostatic mode (BV)
exists and has the group velocity that is antiparallel to the wavevector. Finally, if
the magnetization MS is perpendicular to the film plane, the existing mode is the,
so-called, forward volume magnetostatic mode. (FV) In the latter case the dipole-
dipole matrix element F00 (q) in Eq. (5) can be expressed as:
288 S. O. Demokritov and A. N. Slavin

Fig. 1 Typology of spin wave modes in a magnetic film for different orientations of the
magnetization, MS , and the in-plane wavevector, q

   
F00 q d = P00 q d , (13)

with P00 (q d) given by Eq. (9).

Brillouin Light Scattering a Powerful Tool for Investigation of Spin


Waves

The spin-wave spectrum of a magnetic system can be investigated by various


techniques: ferromagnetic resonance [41], time-resolved Kerr magnetometry
[24, 42–44], and Brillouin light scattering spectroscopy (BLS) [45–47]. The
BLS experimental technique has a number of advantages for the investigations of
magnetic structures. It combines the possibility to study the dynamics of magnetic
systems in the frequency range of up to 500 GHz (corresponding time resolution is
2 ps) with a high lateral resolution of 20–40 μm for the regular setup and down to
220 nm for the, so-called, micro-focus BLS. In both cases the spatial resolution is
defined by the size of the laser beam focus.
Another important advantage of BLS is its very high sensitivity, which allows
us to register thermally excited spin wave modes, so the coherent excitation of
the magnetic element by an external signal is not necessary. This property of the
6 Spin Waves 289

Fig. 2 Scattering process of


a photon from a spin wave hqI hqS
(magnon). θ is the scattering
angle hω I hωS
θ

hq, hω

BLS is especially useful for the experimental investigations of complicated, strongly


confined spin-wave modes in patterned magnetic elements, which will be considered
in the next sections.
The BLS process can be considered as follows (see Fig. 2): monoenergetic
photons (visible light, usually green line 532 nm or blue line 473 nm) with the wave
vector qI and frequency ωI = cqI interact with the elementary quanta of spin waves
(magnons), characterized by the magnon wave vector q and frequency ω. Due to the
conservation laws resulting from the time- and space-translation invariance of a 3D
system the scattered photon increases or decreases its energy and momentum if a
magnon is annihilated or created:

ωS =  (ωI ± ω) (14)

 
q S =  q I ± q (15)

Measuring the frequency shift of the scattered light one obtains the frequency
of the spin wave participating in the BLS process. From Eq. (15) it is evident that
the wave vector qS − qI , transferred in the scattering process, is equal to the wave
vector q of the spin wave. Changing the scattering geometry one can sweep the value
of q and measure the corresponding ω(q). Thus, the spin-wave dispersion ω(q) can
be studied. Note here that for the 3D scattering process the maximum accessible
wavevector q = 2qI , the double value of the light wave vector, corresponds to the
backscattering geometry with the scattering angle θ = 180◦ .
The electromagnetic field of the scattered wave is proportional to the product
of the Faraday/Kerr and other magneto-optical constants of the medium and the
amplitude of the dynamic magnetization, corresponding to the spin wave [45]. Thus,
the BLS intensity, determined by the squared field, is directly proportional to the
dynamic magnetization squared. Magneto-optical effects relate the dielectric tensor
of the medium with Cartesian components of its magnetization. Usually the nondi-
agonal elements of the tensor, the magneto-optical effects (magnetic birefringence
and the Faraday effect and the corresponding dichroisms) are responsible for the
scattering. Therefore, the plane of polarization of the scattered light is rotated by
90◦ with respect to that of the incident light.
The conservation laws, given by Eqs. (14) and (15), follow from the time
invariance of the problem and the translation invariance of an infinite medium,
290 S. O. Demokritov and A. N. Slavin

correspondingly. However, if the scattering volume is finite, the selection rule for
the momentum is broken. For the scattering volume with a size less or comparable
with the wavelength of the light, any spin wave with a wavevector comparable
with that of light contributes to the scattering process. The confinement of the
scattering volume can come from the finite size of the element under consideration
and/or from a small scattering volume of the laser beam. For a thin film or for
nontransparent bulk materials, the thickness of the scattering volume is strongly
confined along the direction normal to the surfaces. Therefore, only the in-plane
wave vector is conserved in the light scattering experiments. As shown in Fig. 3, in
the backscattering geometry the transferred in-plane wavevector q|| is determined
by the angle of incidence q|| = 2qsinα, with q being the absolute value of the wave
vector of the incident light. For green laser light q|| varies in the typical range of
(0–2.5) × 105 cm−1 . This approach is illustrated in Fig. 4 showing the spin-wave
dispersion of a permalloy (Ni80 Fe20 ) film with a thickness of 20 nm, measured in
the DE geometry at the applied magnetic field H = 500 Oe. The experimental data
presented in Fig. 4 were obtained by varying α . The solid line in the figure is the
result of calculation based on Eq. (11) with the value of the permalloy magnetization
4πMS = 9.8 kG.

Fig. 3 Backscattering
process from a thin film. qi is
the wavevector of the incident qs α
light; qs is the wavevector of
the incident light; α is the
angle between the
qi
wavevectors and the film
normal

Fig. 4 The spin-wave


Spin-wave frequency (GHz)

dispersion of a permalloy 14
(Ni80 Fe20 ) film with
parameters listed in the text
12
measured using BLS. The
solid line is the result of
calculation based on Eq. (11) 10

6
0.0 0.5 1.0 1.5 2.0 2.5
5 -1
q|| (10 cm )
6 Spin Waves 291

Spin Waves in 1D Magnetic Elements: Standing and Propagating


Waves

In this section, we will consider spin-wave modes in arrays of micron-size quasi-1D


magnetic elements (stripes and waveguides). We will discuss lateral quantization of
DE modes in a longitudinally magnetized stripe due to its finite width as well as
localization of spin-wave modes near the edges of the stripes. After that, we review
recent experimental investigations of spin-wave propagation, excitation, and control
in microscopic waveguides.

BLS in Laterally Confined Systems

Before we consider new effects resulting from the lateral confinement in stripes, let
us first focus on BLS from laterally confined excitations.
It has been already mentioned in the previous sections that the form of the
conservation laws, which determine the BLS process, depends on the dimensionality
of the studied system: due to lack of translational invariance of a thin magnetic film
along the normal to the film surface, the corresponding component of the wavevector
is not conserved. Instead, the scattering angle (see Fig. 3) determines the 2D in-
plane vector, q|| only, whereas all the thickness modes possessing this q|| contribute
to BLS, albeit with different intensity according to their thickness profiles. If now
the in-plane translational invariance of the magnetic film is broken by patterning,
the in-plane wavevector is no longer fully conserved in the BLS process. In the case
of a spin-wave mode localized in a long stripe, the only conserved component is
the component of along the stripe axis. In analogy with the films, all the laterally
confined modes contribute to BLS, whereas the mode profile along the stripe width
(more specific: the corresponding Fourier component of the dynamic magnetization)
defines the contribution of the mode to BLS. Finally, if the confinement takes
place in all three dimensions, no conservation laws for wavevectors can be applied.
One should perform a Fourier analysis of the 3D distribution of the dynamic
magnetization of a particular mode to calculate its contribution to the BLS intensity.

Lateral Quantization of Spin Waves in Magnetic Stripes

Mathieu et al. [22] and Jorzick et al. [48] investigated spin-wave excitations by BLS
in arrays of permalloy stripes. They have found the effect of lateral quantization of
spin waves due to a finite width of the stripes and observed several dispersionless
spin-wave modes. Since these experiments provide the first account for spin-wave
modes in 1D magnetic systems and heavily contribute to quantitative understanding
of spin wave quantization effects in systems with reduced dimensions, we consider
them in detail.
The samples were made of 20–40 nm thick permalloy films deposited in
UHV onto a Si(111) substrate by means of e-beam evaporation using using
292 S. O. Demokritov and A. N. Slavin

Fig. 5 Scanning electron micrographs of permalloy stripes with a width of 1.8 μm and a
separation of 0.7 μm. (Reprinted with permission from [45], © 2001 by Elsevier)

X-ray lithography with a following lift-off process. Several types of periodic arrays
of stripes with stripe widths w = 1.7 and 1.8 μm and distances between the
stripes above 0.5 μm were prepared. Thus, the interaction between the stripes were
negligible. The length L of the stripes was 500 μm, ensuring 1D-properties of the
stripes. The patterned area was 500 × 500 μm2 , allowing BLS investigation with
a large beam diameter, providing a good wavevector resolution. One of the studied
arrays is shown in Fig. 5. In agreement with the shape-anisotropy arguments, the
magnetic easy axis of the array was along of the stripe axis.
Let us consider a magnetic stripe magnetized in plane along the z-direction
and having a finite width w along the y-direction as shown in Fig. 5. A boundary
condition similar to Eq. (6) at the lateral edges of the stripe should be imposed:

m|x=±d/2 = 0 (16)

One should emphasize that the internal field in the stripe is strongly nonho-
mogeneous due to the nonellipsoidal shape. This nonhomogeneity is of particular
importance close to the edges. In the considered geometry, only the dynamic
internal field is nonhomogeneous, since the static magnetization is along the
edge and does not contribute to the demagnetizing effects. Nevertheless, this
nonhomogeneous dynamic internal field results in specific mechanism for pinning
of the magnetization [39]. Therefore, the Eq. (16) differs from Eq. (6) written for
6 Spin Waves 293

an unconfined film possessing a homogeneous internal field. The corresponding


quantization of qy is then obtained as:


qyn = (17)
w

where n = 1,2, . . . . Using Eqs. (4), (11), and (12) and the quantization expression
(17) one can calculate the frequencies of these so-called width (or laterally
quantized) modes. The profile of the dynamic part of the magnetization m in the
nth mode can be written as follows:


w
mn (y) = An sin y+ (18)
w 2

Equation (18) describes a standing mode consisting of two counter-propagating


waves with quantized wavenumbers, nπ /w. Note here that due to the truncation of
the sin-function at the stripe boundaries the modes are no more infinite plane waves
and the quantized values are not true wavevectors.
In BLS experiments with backscattering geometry the in-plane wavevector q|| ,
transferred in the light scattering process, was oriented perpendicular to the stripes,
and its value was varied by changing the angle of light incidence, α, measured
from the surface normal: as illustrated in Fig. 3. Figure 6 shows a typical BLS
spectrum for the sample with a stripe width of 1.8 μm and as transferred wavevector
q|| = 0.3 × 105 cm−1 , while an external field of 500 Oe was applied along the stripe
axis. As it is seen in Fig. 6, the spectrum contains four distinct modes near 7.8,
9.3, 10.4, and 14.0 GHz. By varying the applied field, the spin wave frequency
for each mode was measured as a function of the field, as displayed in Fig. 7.
The observed dependence of all frequencies on the field confirms that all detected
modes are magnetic excitations. The dispersion law of the modes was obtained by
varying the angle of light incidence, α, and, thus the magnitude of the transferred
wavevector, qy . The obtained results are displayed in Fig. 8 for two with the same
stripe thickness of 40 nm and width of 1.8 pm, but with different stripe separations
of 0.7 pm (open symbols) and 2.2 pm (solid symbols). The dispersion measured
on the arrays with the same lateral layout but with a stripe thickness of 20 nm is
presented in Fig. 9. It is clear from Fig. 8 that one of the detected modes presented
by circles (near 14 GHz) is the PSSW mode, corresponding to p = 1 in Eq. (12).
This mode is not seen in Fig. 9 due to its much higher frequency caused by the
smaller stripe thickness. In the region of low wavevectors the spin wave modes show
a disintegration of the continuous dispersion of the DE mode of an infinite film into
several discrete, resonance-like modes with a frequency spacing between the lowest
lying modes of approximately 0.9 GHz for d = 20 nm and 1.5 GHz for d = 40 nm.
As it is clear from Figs. 8 and 9, there is no significant difference between the
data for the stripes with a separation of 0.7 and 2 μm. This fact indicates that the
mode splitting is purely caused by the quantization of the spin waves in a single
stripe due to its finite width. In other words, the studied stripes can be considered as
independent 1D elements.
294 S. O. Demokritov and A. N. Slavin

PSSW
3000

BLS Intensity (a.u.)


2000

1000

0
6 10 14
Frequency Shift (GHz)

Fig. 6 Experimental BLS spectrum obtained from the stripe array with a stripe thickness of 40 nm,
a width of 1.8 μm, and a separation of 0.7 μm. The applied field is 500 Oe orientated along the
stripe axis. The transferred wavevector of q|| = 0.3 × 105 cm−1 is oriented perpendicular to the
wires. The discrete spin-wave modes are indicated by arrows. PSSW stands for perpendicular
standing spin-wave mode. (Reprinted with permission from [45], © 2001 by Elsevier)

24
Spin Wave Frequency [GHz]

n=3
PSSW n=2
20 n=1

16

12

8 Q-DE

4
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0
Magnetic Field [kOe]

Fig. 7 Frequencies of the in-plane quantized Damon-Eshbach modes (Q-DE) as a function of the
applied field, obtained from the BLS spectra similar to that shown in Fig. 6. The lines are calculated
using Eq. (11) with quantized wavevectors, determined by the quantization numbers n = 1,2,3.
The line labeled PSSW shows the frequency of the first perpendicular standing spin-wave mode.
(Reprinted with permission from [45], © 2001 by Elsevier)
6 Spin Waves 295

17

Spin Wave Frequency (GHz)


16 d = 40 nm
15
14
13
n=5
12
n=4
11
10 n=3

9 n=2
8
n=1
7
6
0.0 0.5 1.0 1.5 2.0 2.5
5 -1
q|| (10 cm )

Fig. 8 Obtained spin-wave dispersion curves for an array of stripes with a stripe thickness of
40 nm, with a width of 1.8 μm and a separation of 0.7 μm (open symbols) and 2.2 μm (solid
symbols). The external field applied along the stripe axis is 500 Oe. The solid horizontal lines
indicate the results of a calculation using Eq. (11) with the quantized wavevectors, determined by
the quantization numbers n = 1,2,3,4,5. The dashed lines showing the hybridized dispersion of the
Damon-Eshbach mode and the first PSSW mode were calculated numerically for a continuous film
with a thickness of 40 nm. (Reprinted with permission from [45], © 2001 by Elsevier)

14
Spin Wave Frequency (GHz)

d = 20 nm
13

12

11

10 n=5
n=4
9 n=3

8 n=2

7 n=1

6
0.0 0.5 1.0 1.5 2.0 2.5
5 -1
q|| (10 cm )

Fig. 9 Obtained spin-wave dispersion curves for an array of stripes for the same conditions as in
Fig. 8, but with a stripe thickness of 20 nm. (Reprinted with permission from [45], © 2001 by
Elsevier)
296 S. O. Demokritov and A. N. Slavin

The main features of the observed spin wave modes in magnetic stripes can be
summarized as follows: (i) For low wavevector values (0–0.8 × 105 cm−1 ) the
discrete modes do not show any noticeable dispersion, behaving like standing wave
resonances. (ii) Each discrete mode is observed over a continuous range of the
transferred wavevector qy . (iii) The lowest two modes appear very close to zero
wavevector, the higher modes appear at higher values. (iv) The frequency splitting
between two neighbored modes is decreasing for increasing mode number. (v)
There is a transition regime (qy = 0.8–1.0 × 105 cm−1 ) where the well-resolved
dispersionless modes converge towards the dispersion of the continuous film (see
dashed lines in Figs. 8 and 9).
All these features can be explained if one implies that the observed discrete,
dispersionless spin wave modes result from a confinement of the DE modes in
the stripes. The confinement causes width-dependent quantization of the in-plane
wavevector of the mode, as discussed above. Considering the corresponding Fourier
component of the dynamic magnetization one can reproduce the measured BLS
intensity of each modes [48].
The frequency of the observed modes can be derived by substituting the
obtained quantized values of the wavevector, qyn , determined by Eq. (17) into the
dispersion equation of the DE mode, Eq. (11). The results of these calculations
are shown in Figs. 8 and 9 by the solid horizontal lines. For the calculation
the geometrical parameters (stripe thickness d = 20 or 40 nm, stripe width
w = 1.8 pm) and the independently measured material parameters 4πMs = 10.2 kG
and γ = 2.95 GHz/kOe were used. Without any fit parameters the calculation
reproduces all mode frequencies very well.

Spin Wave Wells and Edge Modes

In the previous subsection, the simplest geometry with the applied magnetic field
aligned along the stripe axis was considered. In this case, the static field is
homogeneous, while the dynamic field is not homogeneous resulting in the pinned
boundary conditions for the dynamic magnetization. If, however, the applied field is
directed along the width of a thin stripe, both the static and the dynamic internal field
are strongly inhomogeneous. Spin waves propagating along the field are affected in
this case not only by the confinement effects but also by the above inhomogeneities.
Since in unconfined media a wave with q||MS is called the backward volume (BV)
magnetostatic wave, we refer to this experimental geometry as the BV geometry,
contrary to the geometry described in the previous subsection, which we call the
DE geometry. Figure 10 shows two typical BLS spectra obtained from an array of
stripes for the external in-plane magnetic field He = 500 Oe for different orientations
of the field. The spectrum (a) is recorded with both transferred wavevector and He
aligned along the width of the stripe (the BV geometry), whereas the spectrum (b)
is obtained for He oriented along the stripe axis, thus presenting the DE geometry
6 Spin Waves 297

LM

Band
BLS Intensity (a.u.)

(a)

Quantized DE - Modes PSSW

(b)

0 5 10 15 20
Frequency shift (GHz)

Fig. 10 BLS spectra obtained on the stripe array described in the text, q = 0.3 × 105 cm−1 and
He = 500 Oe for (a) the DE geometry and (b) the BV geometry. LM indicates the localized mode.
(Reprinted with permission from [23], © 2002 by the American Physical Society)

discussed in detail above. As it is seen in Fig. 10, both spectra contain several
distinct peaks corresponding to spin-wave modes. The high-frequency peaks can
easily be identified as exchange-dominated PSSW modes. Thus, a narrow PSSW
peak in spectrum (b) confirms the homogeneity of the static internal field in the DE
geometry. On the contrary, a broad PSSW peak in spectrum (a) clearly indicates
a strong inhomogeneous distribution of the internal field across the stripe in the
BV geometry. For a large He the magnetization is parallel to the applied field within
almost the entire stripe. Therefore, poles are created at the edges of the stripe, which
decrease the internal magnetic field in those regions. A detailed analysis shows that
the internal static field, Hi , has a broad maximum in the center of the stripe while it
is vanishing completely near the edges of the stripe [49, 50].
To get further inside into the physics of the observed spin-wave modes, their
dispersion was measured by varying q. It is displayed in Fig. 11 for both orientations
of He . Figure 11a representing the DE geometry is very similar to Figs. 8 and 9: it
clearly demonstrates the lateral quantization of the DE spin waves, resembling a
typical “staircase” dispersion. The frequency of the PSSW mode coincides with
that of the PSSW mode for the unpatterned film and corresponds to an internal
field of Hi = He = 500 Oe, thus corroborating a negligible static demagnetizing
field in the stripes magnetized along their long axes. The dispersion presented in
Fig. 11b and representing the BV geometry differs completely from that shown
in Fig. 11a. First, the PSSW mode is split into two modes, with frequencies
298 S. O. Demokritov and A. N. Slavin

Fig. 11 Spin wave dispersion of the stripe array measured at He = 500 Oe for (a) the DE geometry
with the quantization numbers of the quantized modes as indicated and (b) the BV geometry. In
the latter case, the shadowed region represents the band of non-localized spin modes, whereas LM
indicates the localized mode. The solid lines represent the results of calculation. (Reprinted with
permission from [23], © 2002 by the American Physical Society)

corresponding to internal fields of Hi = 300 Oe and Hi = 0 Oe, respectively, in


agreement with the above qualitative discussion. Second, a broad peak is seen in
the spectra in the frequency range 5.5–7.5 GHz over the entire accessible interval
of q. The shape of the peak varies with q, thus indicating different contributions of
unresolved modes to the scattering cross-section at different q. Third, a separate,
low-frequency, dispersionless mode with a frequency near 4.6 GHz (indicated as
“LM” in Figs. 10 and 11b) is observed over the entire accessible wave vector
range (qmax = 2.5 × 105 cm−1 ) with almost constant intensity. This is a direct
confirmation of a strong lateral localization of the mode within a region with the
width
z = 2π /qmax = 250 nm. From the low frequency of the mode, one can
conclude that it is localized near the edges of the elements, where the internal field
vanishes.
A quantitative analytical description of the spin-wave modes observed in the BV
geometry is made as follows. The frequency of the spin wave as a function of q
and H is given by Eq. (4). Contrary to the DE geometry, where H = He = Hi ,
the demagnetizing effects in the BV geometry are very large: Hi is strongly
inhomogeneous and differs from He . To evaluate Hi :
6 Spin Waves 299

Hi (x, y, z) = He − N (x, y, z) · 4π MS (19)

where Nzz (x, y, z) is the demagnetizing factor. Here we assume a Cartesian


coordinate system, in which the x-axis is perpendicular to the plane of the elements,
the y-axis is along the long axes of the stripes, and the z-axis is along the width of
the stripe: He ||ez .
The field profile of the internal field H(z) obtained from Eq. (19) for He = 500 Oe
is shown in the inset of Fig. 12. For Hi > 0 the magnetization is parallel to He .
Near the edges, however, regions with Hi = 0 and with continuously rotating
magnetization are formed [49, 50], which reflect spin waves propagating from the
middle of the stripe towards these regions. Moreover, a spin wave propagating
in an inhomogeneous field might encounter the second turning point even if the
magnetization is uniform. In fact, for large enough values of the internal field there
are no allowed real values of q, consistent with the spin-wave dispersion (Eq. 4) – a
potential well for propagating spin waves is created. Similar to the potential well in
quantum mechanics, the conditions determining the frequencies fr of possible spin-
wave states in the well created by the inhomogeneous internal field are as follows:


2 q (H (z), f ) dz +
ψ1 +
ψ2 = 2rπ (20)

where r = 1,2,3, . . . and q(H(z),f ) is found from the spin-wave dispersion Eqs. (4)
and (5), and
ψ l ,
ψ 2 are the phase jumps at the left and right turning points,
between which Eq. (4) has a real solution q(z) for a fixed frequency f.
We will illustrate these ideas in the following. The dispersion curves for spin
waves with q||He and p = 0 calculated using Eqs. (4) and (5) for different constant
values of the field are presented in Fig. 12. A dashed horizontal line shows the
frequency of the lowest spin-wave mode f1 = 4.5 GHz obtained from Eq. (20) for
the lowest value r = 1 in good agreement with the experiment. It can be seen from
Fig. 12 that for H > 237 Oe there are no spin waves with the frequency f1 = 4.5 GHz.
Therefore, the lowest mode can exist only in the spatial regions in the magnetic
stripe where 0 Oe < H < 237 Oe. The corresponding turning points are indicated in
the inset of Fig. 12 by the vertical dashed lines. Thus, the lowest mode is localized
in the narrow region
z near the lateral edges of the stripe where 0.26 < |z/w| < 0.39.
The mode is composed of exchange-dominated plane waves with qmin < q < qmax ,
as indicated in Fig. 12.
The higher-order spin-wave modes with r > 1 having their frequencies above
5.3 GHz are not strongly localized under the used experimental conditions and exist
everywhere in the stripe where the internal field is positive (0 < |z/w| < 0.39). In
the experiment, they show a band, since the frequency difference between the fr and
fr + 1 modes is below the frequency resolution of the BLS technique. Note here that
several localized modes can be observed at higher values of He [51].
300 S. O. Demokritov and A. N. Slavin

Fig. 12 Dispersion of plane spin waves in the BV geometry at constant internal fields as indicated.
Inset: the profile of the internal field in a stripe.
z shows the region of the lowest mode
localization. (Reprinted with permission from [23], © 2002 by the American Physical Society)

Implementation of Micro-Focus BLS for Laterally Patterned


Magnetic Systems

All the above BLS data were obtained in the so-called Fourier microscope mode
[48]. In this case, the diameter of the laser beam was kept quite large (typically
30–50 μm), allowing fulfilling the wavevector conservation law Eq. (15), and the
frequencies and the intensities of the studied spin-wave modes were investigated as
a function of transferred wavevector.
A complimentary approach is the micro-focus BLS [46]. Here the coherent laser
light focused onto the surface of the magnetic system into a diffraction-limited spot
by using a high-quality microscope objective lens with a large numerical aperture.
While the frequency shift of the scattered light is equal to the frequency of the
magnetization oscillations, its intensity (referred to as BLS intensity) is proportional
to the intensity of magnetization oscillations at the position of the probing spot.
The latter fact enables direct spatial imaging of the spin-wave intensity by two-
dimensional rastering of the probing spot over the sample surface (Fig. 13a). The
acquired intensity maps, such as that shown in Fig. 13b, allow one to obtain
information about the spatial characteristics of spin waves. In the case of spin-wave
6 Spin Waves 301

a)
Microwave
current

Probing
Antenna
light

H0 Magnonic
waveguide

Substrate

b) 1 BLS c)
intensity,
-1
a.u.
10

-2
10 1

cos (ϕ) 0
500 nm
-1

Fig. 13 (a) Sketch of a typical micro-focus BLS experiment on spin-wave propagation in a


microscopic waveguide. (b) and (c) Representative examples of the two-dimensional maps of the
spin-wave intensity (b) and phase (c) recorded by micro-focus BLS. (© 2015 IEEE. Reprinted,
with permission, from [47])

beams propagating in waveguides, they also provide important information about


the damping and the spatial characteristics of the spin-wave beam.
For reliable two-dimensional imaging of spin waves, the spatial resolution of the
micro-focus BLS apparatus is of crucial importance, which is found to be about
250 nm for the wavelength of 532 nm in agreement with classical optics. This
resolution can be further improved to about 50 nm [52] by utilizing the principles of
near-field optical microscopy. However, the use of this approach inevitably leads to
a reduction of the sensitivity, which noticeably increases the time needed for BLS
measurements. Therefore, the experiments described here were performed by using
free-space-optics micro-focus BLS apparatus with the resolution of 250 nm, which
is sufficient to address most of the spin-wave propagation phenomena.
In agreement with the discussion in Sect. “BLS in Laterally Confined Systems”,
high-spatial resolution of the micro-focus BLS technique is incompatible with the
wavevector resolution due to the large uncertainty of the light-scattering angle
associated with the tight focusing of the probing light. Thus, the information about
the wavevector of the wavelength of the studied spin waves is lost, and the spin-
wave dispersion cannot be obtained on the usual way. However, this drawback can
be eliminated by making use of the time invariance of the BLS-process, which
results in frequency/phase conservation in the light-scattering process. It means
302 S. O. Demokritov and A. N. Slavin

that the phase of the scattered light is directly correlated to the phase of the
magnetization oscillations in the spin-wave mode. This phase information can be
acquired by utilizing interference of the scattered light with the light modulated
by the signal used to excite the magnetization oscillations. This approach enables
direct measurements of the phase difference between the excitation signal and
the phase of a propagating spin wave at the given spatial location, providing, for
example, direct information about the spin-wave wavelength. The phase-resolution
technique was first demonstrated for standard macro-BLS apparatus [53] and was
subsequently adapted for micro-focus BLS measurements [54, 55]. Figure 13c
shows a representative example of the measured spatial phase map for spin waves
propagating in a submicrometer-width magnonic waveguide, excited by microwave
current in the antenna. The plotted value is cos(ϕ), where ϕ is the phase difference
between the microwave current and the magnetization oscillations in the spin wave.
The spatial period of cos(ϕ)-function is equal to the spin-wave wavelength at a given
excitation frequency. Therefore, by varying the latter and measuring the spatial
period, one can obtain the complete information about the spin-wave dispersion
characteristics of the studied waveguide.

Propagating Waves in 1D Magnetic Structures

By analyzing the lateral quantization of spin waves in stripes in the previous sub-
sections, we have considered standing waves, i.e., we implied that the component of
the wavevector along the stripe axis is zero. To extend this approach to propagating
waves in such a waveguide we should consider two components of the wavevector:
one component quantized due to finite width of the stripe and another component
continuously varying as illustrated in Fig. 14. In fact, if the quantized components
qzn are known, the spectrum of normal waveguide modes can be obtained from the
two-dimensional dispersion surface described by Eq. (4) and by cutting it along qy
at the fixed qzn as illustrated in Fig. 14a by the curves labeled as DE1 − DE3 . For
the sake of clearness we project these curves onto the frequency- qy plane, as shown
in Fig. 14b, keeping in mind that the different curves correspond to different qzn .
As seen from Fig. 14b, the considered modes propagate perpendicular to He , i.e.,
they are analogues of the DE mode in an extended film, their dispersion curves are
shifted to lower frequencies with respect to that of the unconfined DE mode, and this
shift increases with the increase of the mode number. This is not surprising, since all
these modes are characterized by a nonzero component of the wavevector qz parallel
to He . Since the dipolar magnetic energy is known to decrease with the increase
of this component causing the backward dispersion of the BV modes (Fig. 14a),
the dispersion curves of the waveguide modes shift to lower frequencies with the
increase of the mode number, which corresponds to the increase in qz .
We emphasize that the above-described approach to calculation of the dispersion
curves of the normal waveguide modes is a rough approximation, since they do not
take into account the reduction of the static magnetic field inside the waveguide
6 Spin Waves 303

Fig. 14 (a) Two-dimensional a)

Frequency
z DE
dispersion spectrum of spin
waves in an extended in-plane y DE1
magnetized ferromagnetic
film. Inset shows the DE3
geometry of the stripe DE2 f0 DE2
waveguide and transverse DE1 BV
profiles of the dynamic
magnetization for normal DE3 qy
waveguide modes. (b)
Calculated (solid lines) and H0
qz
measured (symbols)
dispersion curves for a b)
waveguide with the width
10
DE1
w = 800 nm and the Frequency, GHz DE
thickness d = 20 nm DE2
f1
magnetized by the static field 9 DE3
He = 900 Oe. Dashed line
shows the dispersion curve
for Damon-Eshbach mode in 8
an extended film. (© 2015 q1 q2 q3
IEEE. Reprinted, with 0 1 2 3 4 5 6
permission, from [47]) -1
qy, μm

caused by the demagnetization effects, which can be further taken in account using
Eq. (19).
The data of Fig. 14b show that the dispersion spectrum of waveguide modes
supports multimode propagation of spin waves at all frequencies above f0 . For exam-
ple, by exciting spin waves at the frequency f1 (see Fig. 14b), one simultaneously
excites a number of modes with different longitudinal wavevectors qy . Neglecting
attenuation of spin waves, the spatial distribution of the intensity of the dynamic
magnetization in these patterns can be described as (cf. Eq. 18):
2


w  

I (y, z) = An sin z+ exp −iq n y (21)
n w 2

where An are the amplitudes of the modes and qn are their longitudinal wavevectors
at the given excitation frequency (see Fig. 14b). Figure 15 shows the results of
calculations based on Eq. (21) performed for different ratios between the amplitudes
An and the dispersion data taken from Fig. 14b. In the simplest case, where the
only present mode is the fundamental mode with n = 1 (Fig. 15a), the intensity
distribution is uniform in the longitudinal direction and shows a half-sine profile
in the transverse direction. The co-propagation of the fundamental mode and the
mode with n = 2 (Fig. 15b) results in an appearance of a “snake”-like pattern,
which becomes more pronounced with the increase of A2 . We note that, because
of the symmetry reasons, the mode n = 2 possessing an antisymmetric distribution
of the dynamic magnetization across the waveguide width (inset in Fig. 14a) is
304 S. O. Demokritov and A. N. Slavin

Fig. 15 Interference patterns a) A1=1, A2=0, A3=0


for the three lowest-order 0.4

z, μm
waveguide modes calculated 0
for different ratios between
their amplitudes, as labeled. -0.4
Calculations were performed b) A1=1, A2=0.15, A3=0
for the waveguide with the 0.4

z, μm
width of 800 nm and the 0
thickness of 20 nm
magnetized by the field -0.4
A1=1, A2=0.3, A3=0
He = 900 Oe. (© 2015 IEEE. 0.4
z, μm
Reprinted, with permission,
0
from [47])
-0.4
c)
A1=1, A2=0, A3=0.15
0.4
z, μm

0
-0.4
A1=1, A2=0, A3=0.3
0.4
z, μm

0
-0.4
0 1 2 3 4 5
y, μm

normally not excited in axially symmetric guiding systems and can only be observed
if this symmetry is broken. In contrast, a significant contribution of the mode
with n = 3 is detected in most of the experiments. As seen from Fig. 15c, the
co-propagation of this mode and the fundamental mode of the waveguide results
in a periodic spatial beating pattern, where the spin-wave energy is periodically
concentrated in the middle of the waveguide, while the transverse width of the spin-
wave beam shows a periodic modulation. By analogy with the light focusing in
optics, this effect was given a name of “spin-wave focusing” [56]. Figure 16a shows
a typical measured spin-wave intensity map for a 2.4 μm wide and 36 nm thick
Py waveguide clearly demonstrating this effect (compare with Fig. 15c). In order
to highlight the details of the interference pattern in Fig. 16a, the spatial decay
of spin waves is numerically compensated by multiplying the experimental data
by exp.(2y/ξ), where ξ is the spin-wave decay length – the distance over which
the wave amplitude decreases by a factor of e. The latter is determined from the
dependence of the BLS intensity integrated across the transverse waveguide section
versus the propagation coordinate (solid symbols Fig. 16b). As seen from Fig. 16b,
spin waves in the studied waveguide exhibit clear exponential decay characterized
by ξ = 6.4 μm. Figure 16b also shows by open symbols the transverse width
of the spin-wave beam versus the propagation coordinate, which can be used to
quantitatively characterize the strength of the focusing effect. In particular, for the
data of Fig. 16b, the modulation of the beam width caused by the focusing is equal to
about 70%, while the smallest width observed at the focal point is equal to 0.65 μm.
6 Spin Waves 305

a) Propagation

BLS intensity, a.u.


1 1

z, μm
0 0.5

-1 0
0 1 2 3 4 5 6 7
y, μm
b) 1 1.5

Beam width, μm
intensity, a.u.
Integral

1.0

0.1 0.5
0 1 2 3 4 5 6 7
y, μm

Fig. 16 (a) Measured map of the spin-wave intensity for a waveguide with the width of 2.4 μm
and the thickness of 36 nm magnetized by the field of 900 Oe. Excitation frequency is 9.4 GHz.
Spatial decay of spin waves is numerically compensated. (b) Solid symbols – BLS intensity
integrated across the transverse waveguide section versus the propagation coordinate in the log-
linear scale. Line is the exponential fit to the experimental data. Open symbols – transverse width
of the spin-wave beam measured at one half of the maximum intensity versus the propagation
coordinate. (© 2015 IEEE. Reprinted, with permission, from [47])

In the above discussion on the normal waveguide modes, we neglected the


nonuniformity of the magnetic field inside the waveguide caused by the demag-
netization effects. On one side, this nonuniformity does not qualitatively affect
the structure of the modes evolving from plane spin waves due their geometrical
confinement. As discussed above, the nonuniformity is known [49] to result in the
appearance of the regions of strongly reduced internal field close to the edges of the
stripe [50], which gives rise to additional spin-wave modes having no analogue in
the case of extended magnetic films [23]. According to [49], the distribution of the
internal field across the width of a magnetic stripe magnetized perpendicular to its
axis can be approximated as (cf. Eq. 18):
    
4π MS d d
Hi (z) = He − atan − atan (22)
π 2z + w 2z − w

Figure 17a shows this distribution calculated for the waveguide with the width
of w = 2.1 μm and the thickness of d = 20 nm magnetized by the static field
He = 1100 Oe. As seen from this data, close to the edges of the waveguide, the
internal field is drastically reduced resulting in the appearance of field-induced
channels, where spin waves can be localized, as discussed in previous subsection.
306 S. O. Demokritov and A. N. Slavin

a)
1000

field, Oe
Internal
500 Field-induced
channels
0
-1.0 -0.5 0.0 0.5 1.0
z, μm

b) 9.8 GHz Propagation


1
z, μm

BLS intensity, a.u.


1
-1
9.0 GHz 0.5
1
z, μm

0 0
-1
0 2 4 6 8 10
y, μm
c) 1.5 0.6
Distance between

Beam width, μm
the beams, μm

1.0 0.5

0.5 0.4
9.0 9.2 9.4 9.6 9.8
Frequency, GHz

Fig. 17 (a) Calculated distribution of the internal static magnetic field across the width of a
waveguide with the width of 2.1 μm and the thickness of 20 nm magnetized by the static field of
1100 Oe. Horizontal dashed line marks the value of the external magnetic field. (b) Measured maps
of the spin-wave intensity for two excitation frequencies, as labeled. Spatial decay of spin waves
is numerically compensated. (c) Distance between the centers of the spin-wave beams and their
transverse width measured at one half of the maximum intensity versus the spin-wave frequency.
(© 2015 IEEE. Reprinted, with permission, from [47])

Since these modes are mostly concentrated in the areas of the reduced field,
their typical frequencies are lower than the frequencies of the “center” modes as
discussed above. Therefore, in order to address them experimentally, one needs
to excite the waveguide at frequencies below the frequency of the uniform ferro-
magnetic resonance f0 (see Fig. 14a). Figure 17b shows two decay-compensated
spin-wave intensity maps measured for a waveguide with the parameters given
above by applying the excitation signal at the frequencies of 9.0 and 9.8 GHz, which
are smaller than f0 equal to about 10 GHz for the used experimental conditions.
6 Spin Waves 307

The data of Fig. 17b show that, in agreement with the simple qualitative model,
at these frequencies, the spin waves do not occupy the entire cross-section of
the waveguide. Instead, they form two narrow beams with the submicrometer
width whose spatial positions depend on the spin-wave frequency. The quantitative
analysis (see Fig. 17c) shows that, in the wide frequency interval, the widths of
the beams vary moderately staying in the range 400–500 nm, while the distance
between their centers monotonously increases with the decrease of the frequency
from 0.8 μm to 1.4 μm, i.e., by more than 70%.

Control and Conversion of the Propagating Waves

One of the great advantages of spin waves for implementation of signal-processing


devices is their controllability by the static magnetic field, which allows one to
efficiently manipulate the spin-wave propagation. Although this control mechanism
is straightforward, its implementation on the macroscopic scale requires the use of
electromagnets making the resulting devices extremely space and power consuming.
The downscaling of spin-wave devices provides a route for overcoming this
drawback, since, in microscopic systems, the control magnetic field has to be created
over small distances and sufficiently large local magnetic fields can be created
by using relatively small electric currents [55]. This approach is schematically
illustrated in Fig. 18a. Instead of using external electromagnets, the control magnetic
field is created by the electric current flowing in a control line, which is directly
integrated into the waveguide. The composite waveguide consists of two layers:
the upper 20 nm thick Py layer guiding spin waves and the bottom 100 nm thick
Cu layer used as a current-carrying line to generate controlling magnetic fields.
Because of the large difference in conductivities of Cu and Py, the shunting of the
control current through the Py layer is negligible, which makes electrical isolation
between the two layers unnecessary.
The Oersted field in the Py layer produced by the current in the Cu layer can be
approximated as
H = I/(2w + 2d), where I is the current strength and w = 0.8 μm
and d = 0.1 μm are the width and the thickness of the Cu line, respectively. Due to
the small cross-section of the control line, one achieves sufficiently high efficiency
of the magnetic field generation of about 7 Oe/mA. The effect of the current on the
dispersion characteristics of spin waves in the waveguide is illustrated by Fig. 18b.
By applying I = ± 12 mA one creates controlling magnetic field
H = ± 83 Oe,
which adds or subtracts from the static magnetic field He = 520 Oe resulting in
the shift of the dispersion curve by more than ±500 MHz. This shift leads to
a significant variation of the longitudinal wavevector qy at the given spin-wave
frequency. Figure 18c demonstrates the controllability of the wavevector and the
wavelength of sin waves with the frequency of 7 GHz. These data show that by
applying I = ±12 mA one can change these parameters by more than 50%. Note
that the variation of qy with the current is nearly linear. Since the phase accumulated
by the spin wave over a propagation distance L is proportional to qy :
ϕ = Lqy , this
also implies a linear controllability of the phase accumulation, which is attractive
308 S. O. Demokritov and A. N. Slavin

Fig. 18 (a) Schematic of a z


a)
composite waveguide with y
the integrated control Cu line. He Py
(b) Dispersion curves of the (20 nm)
fundamental waveguide mode Cu
for different currents in the (100 nm)
control line, as labeled. Lines dc DH
show the calculated
dispersion curves, and b) 8
symbols show the results of I=12 mA

Frequency, GHz
measurements by 7
phase-resolved BLS
technique. (c) Longitudinal 6
wavevector qy and the I=-12 mA
wavelength of the spin wave 5 I=0
at the frequency of 7 GHz
versus the control current. 0 1 2 3 4 5 6
Symbols – experimental data, -1
qy, μm
lines – guides for the eye. (©
2015 IEEE. Reprinted, with c) 5 2.5
permission, from [47])

Wavelength, μm
-1

2.0
qy, μm

1.5
3
1.0
-15 -10 -5 0 5 10 15
Current, mA

for technical applications. Based on the data of Fig. 18c, one can conclude that by
applying the control current of ±12 mA, the phase accumulated by the spin wave
can be changed by ±π radians at the propagation distance of about 3.2 μm which is
smaller than the spin-wave propagation length. This makes the proposed mechanism
well suited for implementation of magnonic logic devices [57], where the digital
information is coded into the phase of propagating spin waves.
In addition to the use of magnetic fields created by electric currents, the control
of spin-wave propagation can also be realized by using demagnetizing fields. Since
the demagnetizing field depends on the ratio between the width and the thickness of
the waveguide (Eq. 22), simple variation of one of these parameters enables efficient
manipulation of spin waves [58, 59].
A waveguide with the varying width is schematically shown in Fig. 19a: while
the thickness of the waveguide d = 36 nm remains constant, its width w varies from
1.3 to 2.4 μm over a transition region with the length L. According to Eq. (22),
such a variation results in a spatial variation of the internal field, which changes
from 870 Oe in the narrow part of the waveguide to 950 Oe in the wide part
(Fig. 19b). Due to this variation, the dispersion curves of the fundamental center
waveguide mode are shifted in the two parts by about 500 MHz in the frequency
domain (Fig. 19c). If the excitation frequency is chosen to be located between the
6 Spin Waves 309

a) z b)

waveguide
y 1.3 μm 950

Narrow
field, Oe

waveguide
Internal
He

Wide
900
L
850
2.4 μm
0 1 2 3 4 5 6
y, μm
c) 11
Frequency, GHz

Wide
f2 waveguide
10

f1 Narrow
9 waveguide

0 1 2 3
-1
qy, μm
d)
Propagation
8.7 GHz 9.1 GHz

1 μm
e) L=3 μm L=1 μm

Fig. 19 (a) Schematic of a spin-wave waveguide with a varying width. (b) Calculated distribution
of the internal static magnetic field in the section along the axis of the waveguide with the
thickness of 36 nm and the geometrical parameters given in (a). The external static magnetic field
He = 1000 Oe. (c) Calculated dispersion curves for the fundamental center waveguide mode in the
wide and the narrow parts of the waveguide. (d) Maps of the spin-wave intensity measured at the
excitation frequencies of 8.7 and 9.1 GHz, as labeled. Width of the transition region L = 2 μm. (e)
Maps of the spin-wave intensity measured at the excitation frequency of 9.7 GHz in waveguides
with L = 3 and 1 μm, as labeled. In (d) and (e) the spatial decay of spin waves is numerically
compensated. (© 2015 IEEE. Reprinted, with permission, from [47])

cut-off frequencies of the two dispersion curves, i.e., between 8.7 and 9.2 GHz (f1 in
Fig. 19c), the center mode propagating in the narrow part of the waveguide cannot
pass into the wide part. Instead, it should be transformed into the edge mode, whose
frequency range is located below that of the center modes. This case is illustrated
in Fig. 19d showing two spin-wave intensity maps measured for the excitation
frequencies of 8.7 and 9.1 GHz. These maps clearly demonstrate the conversion
of the center mode into the edge mode characterized by two narrow spin-wave
beams with frequency-dependent spatial separation (see Sect. “Implementation of
Micro-Focus BLS for Laterally Patterned Magnetic Systems”). Since the two beams
310 S. O. Demokritov and A. N. Slavin

propagate in the field-induced channels and are independent from each other, the
observed transformation can be used for implementation of a spin-wave splitter.
One also observes interesting behaviors in the case, when the frequency of
spin waves is larger than cut-off frequencies in both parts of the waveguide (f2 in
Fig. 19c). In the waveguides with the relatively long transition region (L = 3 μm
in Fig. 19d) the propagation of spin waves from the narrow to the wide part
is quasi-adiabatic. It is only accompanied by the increase in the wavelength,
while the spatial structure of the spin-wave beam remains unchanged. However,
in systems with shorter transitions (e.g., L = 1 μm in Fig. 19e), the propagation
is accompanied by an appearance of a complex intensity pattern, which can be
recognized as an interference pattern created by several waveguide modes with
comparable amplitudes (see Fig. 15c). This is due to the strong coupling of the
waveguide modes mediated by the spatial nonuniformity in the waveguide, which
results in the efficient energy transfer from the fundamental mode to the higher-
order modes. As seen from Fig. 19e, this effect causes a strong concentration of
the spin-wave energy in the middle of the waveguide at a certain distance from the
transition region, which can be treated as an enhanced spin-wave focusing.
A particularly interesting case is a junction between a 1D waveguide and 2D
film [60, 61] (Fig. 20a). Because of the abrupt transition in such a system, the
wavevector of spin waves is not conserved during the conversion of the waveguide
modes into the modes of the extended film. In other words, being radiated from
the open end of the waveguide, the waveguide mode excites spin waves within
a large range of wavevectors. Because of the temporal translation symmetry, the
frequency of radiated spin waves is equal to that of the waveguide mode. Therefore,
the characteristics of the radiated waves can be obtained by considering constant-
frequency contours of the two-dimensional dispersion surface (Fig. 14a), as shown
in Fig. 20b. Figure 20c shows such contours projected onto the qz − qy plane,
calculated for the conditions used in the experiment: d = 36 nm, He = 690 Oe.
In this representation the vector of the group velocity of spin waves Vg is directed
along the normal to the constant-frequency contour: Vg = 2π ∇ f (qy , qz ). As
seen from Fig. 20c, except for the region of small qz , the direction of the group
velocity is practically constant and builds a well-defined angle with the direction
of the static magnetic field He . This indicates that a large group of spin waves with
different wavevectors transmits energy in the same direction. Therefore, one expects
predominant radiation of the spin-wave energy from the waveguide along this
direction. Figure 20d shows an experimental spin-wave intensity map corresponding
to the frequency of 8.2 GHz. The shown experimental data confirm the conclusions
of the above analysis: the spin waves are radiated in a form of two narrow beams and
their directions coincide well with the direction of the group velocity obtained from
the analysis of the 2D dispersion surface (arrows in Fig. 20d). Since the direction
of the beams is determined by He , it can be electronically steered. In general, the
phenomenon enables a directional, 1D transmission of spin waves in an extended
2D magnetic film without utilization of the geometrical confinement. Recently, this
phenomenon was also observed for spin waves radiated by a spin-torque nano-
oscillator [62], which, due to its small size, also emits spin waves within large
interval of wavevectors.
6 Spin Waves 311

a) b)

Frequency
z DE
y
f2
Waveguide

He f0 f1

BV f0

qy
Extended film
qz

8.2 GHz 10.2 GHz


c) d)
4
Waveguide
2 Vg
-1
qz, μm

-2 He

-4
0 2 4 6 1 μm
-1
qy, μm

Fig. 20 (a) Schematic of a junction between 1D waveguide and 2D extended film. (b) Two-
dimensional dispersion surface of spin waves in an extended magnetic film with marked constant-
frequency contours. (c) Constant-frequency contours projected onto the qz -qy plane calculated for
t = 36 nm and He = 690 Oe. (d) Measured map of the intensity of spin waves with the frequency
of 8.2 GHz radiated from a waveguide with the width of 2 μm into an extended magnetic film.
The spatial decay of spin waves is numerically compensated. (© 2015 IEEE. Reprinted, with
permission, from [47])

Inductive Excitation of Spin Waves in 1D Waveguides

In spite of the recent progress in studies of spin-transfer torque excitation of


propagating spin waves (see Sect. “Spin Waves in 0D”), the inductive excita-
tion mechanism shortly introduced above still remains the most widely used in
experimental investigations of spin-wave phenomena in both 2D and 1D magnetic
systems, because its implementation does not require complex nanolithography
techniques. This method is also characterized by the full control over the frequency
of excited spin waves, which makes it attractive for research purposes. The inductive
excitation by means of spin-wave antennae was widely used in the past for imple-
mentation of macroscopic-scale devices (see, e.g., [63, 64]) and was subsequently
transferred onto the microscopic scale without significant modifications.
Figure 21a shows schematics of a spin-wave waveguide with a spin-wave antenna
on top. A microwave-frequency electric current transmitted through the antenna
creates the dynamic magnetic field h, which couples to the dynamic magnetization
in the waveguide and excites propagating spin waves. Experimentally, the excitation
312 S. O. Demokritov and A. N. Slavin

process can be efficiently characterized by micro-focus BLS by placing the probing


laser spot onto the waveguide in the vicinity of the antenna and recording the BLS
intensity as a function of the frequency of the excitation current. Such an excitation
curve is shown in Fig. 21b. The curve exhibits a fast drop of the spin-wave intensity
at frequencies below the frequency of the ferromagnetic resonance f0 . However, the
intensity of excited spin waves still remains noticeable in this region and shows a tail
extending far into the low-frequency spectral interval. Within this frequency region,
propagating edge modes are excited. The part of the excitation curve at frequencies
above f0 corresponding to the center waveguide modes shows a nonmonotonous
behavior with several oscillations. This oscillatory behavior can be understood
based on the spin-wave excitation theory adapted for the case of microscopic
waveguides [65]. According to this theory, the amplitudes of the waveguide modes
An excited by the antenna are determined by the spatial overlap of the dynamic
magnetic field h(x,y,z) created by the antenna with the dynamic magnetization in
the waveguide m(x,y,z). As seen from Fig. 21a, the former has x- and y-components.
Both these components are perpendicular to the direction of the static magnetization
and, therefore, can linearly couple to the dynamic magnetization. However, because
of the strong dynamic demagnetization in the thin-film waveguides, the effect of
the out-of-plane component hx is relatively small and can be neglected in the first
approximation. Then the excitation problem is reduced to the consideration of
the spatial overlap of the hy component with the corresponding component of the
magnetization. Neglecting the variations of the field and the dynamic magnetization
across the waveguide thickness, the amplitudes of excited spin-wave modes can be
expressed as:


w/2 ∞


An ∝ n
hy (z)my (z)dz · hy (y)my (y)dy
n
(23)

−w/2 −∞

where w is the width of the waveguide. The corresponding profiles of components


of the field and the dynamic magnetization are schematically shown in the inset
in Fig. 21a: hy (z) = const, and hy (y) can be approximated by a rectangular
pulse function with the width equal to that of the antenna d. As discussed in
Sect. “Propagating Waves in 1D Magnetic Structures” (Eq. 21), the transverse
profiles of the dynamic magnetization corresponding to the center modes can
approximated as mny (z) ∝ sin (nπ (z + w/2)), while the longitudinal profile


represents a propagating wave: mny (y) ∝ exp −iq ny y , where qyn are the
longitudinal wavevectors of the modes at the given spin-wave frequency.
Substituting these expressions into Eq. (22) one obtains:



nb
1 − (−1) sin
n q y
An ∝ (24)
n qyn
6 Spin Waves 313

x Antenna
a)
Microwave z hy(y)
current y
b

h
He
my(y)
Magnetic
waveguide y

n=1 n=2 n=3


hy(z)
n
my (z) z

b)
Edge DE
BLS Intensity,

Modes Modes
a.u.

f0

c) 8
n=3
6
-1

2π/d
qy, μm

n=1
4
2
0

d)
1.0
n=1
An, a.u.

0.5
n=3

0.0
7 8 9 10 11 12 13
Frequency, GHz

Fig. 21 (a) Schematic of a waveguide with the spin-wave antenna on top. Inset schematically
shows the y- and z-profiles of the dynamic field of the antenna and those of the dynamic
magnetization in the spin-wave modes. (b) Spin-wave excitation curve measured in a waveguide
with the width of 2 μm and the thickness of 36 nm magnetized by the static field of 900 Oe. (c)
Calculated dispersion curves for the first two symmetric waveguide modes. Horizontal dashed line
shows the value of qy , at which the excitation efficiency vanishes. Arrows show the corresponding
frequencies for the two modes. (d) Calculated amplitudes of the first two symmetric waveguide
modes versus the excitation frequency. The vertical dashed line in (b)–(d) marks the frequency of
the uniform ferromagnetic resonance f0 . (© 2015 IEEE. Reprinted, with permission, from [47])
314 S. O. Demokritov and A. N. Slavin

The first term in Eq. (24) shows that the amplitudes of the modes decrease
with the increase of the mode number as 1/n and that only modes with symmetric
transverse profiles can be excited. The second term shows that the excitation
efficiency has a maximum at qy = 0, exhibits an oscillatory behavior in agreement
with the experimental data of Fig. 21a, and vanishes for qy = 2π m/b or b = mλ,
where m = 1,2,3 . . . and λ is the wavelength of the spin wave.
Figure 21d presents the frequency dependences of the amplitudes of the first
two symmetric waveguide modes calculated based on Eq. (23) and the dispersion
curves obtained by using Eqs. (4) and (5) (shown in Fig. 21c). The calculations
were performed for the conditions used to acquire the experimental excitation curve
shown in Fig. 21b: w = 2 μm, d = 36 nm, He = 900 Oe, and b = 1.5 μm. As
seen from Fig. 21d, the fundamental mode with n = 1 clearly dominates over
the mode with n = 3 and the nodes of the corresponding curve match well with
those seen in the experimental data. We note that, due to the shift of the dispersion
curves corresponding to different modes (Fig. 21c), the frequencies, at which the
excitation efficiency vanishes, are different for the modes with n = 1 and 3. Since
the contribution of the mode with n = 3 is relatively small, vanishings of its
amplitude cannot be seen in the experimental curve in Fig. 21b. Nevertheless, this
vanishing can be proven by the spatial imaging of the spin-wave propagation. These
measurements show a strong reduction in the spatial transverse modulation of the
spin-wave beam at the frequency of about 10 GHz in agreement with the data
of Fig. 21d. We emphasize that the dependence of the relative amplitudes of the
waveguide modes on the excitation frequency allows one to control the strength of
the spin-wave focusing effect. If the transverse modulation of the spin-wave beam
is not desired, one can choose the spin-wave frequency in the vicinity of the point
of the vanishing excitation efficiency of the mode with n = 3, while by choosing
the frequency in the region, where the excitation efficiencies of both modes are
approximately equal, one obtains the strongest mode interference.
The above-described simple model can be extended by additionally taking
into account the out-of-plane component of the dynamic magnetic field of the
antenna [65]. This extension does not significantly modify the discussed frequency
dependence of the excitation efficiency. However, it results in different excitation
efficiencies of spin waves propagating in the positive and the negative direction of
the y-axis. This excitation non-reciprocity is often confused with the intrinsic non-
reciprocity of DE modes [19], which has no effect for spin waves with qy < <1/d
addressed in most of the experiments utilizing the inductive spin-wave excitation.
For typical parameters of the Py micro-waveguides, the excitation non-reciprocity
results in the factor of 4–5 difference between the intensities of waves propagating
in the opposite directions from the antenna.
The above analysis shows that by using an inductive spin-wave antenna one
can only efficiently excite spin waves within a certain interval of wavevectors
corresponding to the first oscillation lobe of the excitation efficiency function
(Fig. 21d), whose spectral width is determined by the width of the spin-wave
antenna b. This limits the applicability of standard inductive excitation, since, in
order efficiently to excite short-wavelength spin waves necessary for most of the
magnonic applications, one has to reduce the width of the antenna, which inevitably
6 Spin Waves 315

causes the microwave impedance matching problems and increases the microwave
losses in magnonic devices. This drawback can be partially overcome [59] by
utilizing the spin-wave controllability by the demagnetizing fields discussed in
Sect. “Propagating Waves in 1D Magnetic Structures”.
Figure 22a schematically shows a system enabling inductive excitation of spin
waves with the wavelength smaller than the width of the antenna: the width of
the waveguide equal to 2 μm in the excitation area gradually reduces to 0.5 μm,
while the thickness of the waveguide d = 40 nm stays unchanged. As discussed
in detail in Sect. “Propagating Waves in 1D Magnetic Structures”, the variation of
the waveguide width leads to the variation of the internal field along the waveguide
axis, which causes the shift of the dispersion curves of the waveguide modes in
the frequency domain. As a result, spin waves excited in the 2 μm-wide part of
the waveguide continuously decrease their wavelength propagating in the tapered
part. This process is illustrated by the two-dimensional phase map (Fig. 22b) and its
section along the waveguide axis (Fig. 22c) obtained by using the phase-resolved
micro-focus BLS technique. The experimental data clearly show a significant
reduction of the wavelength, which, after passing the tapered part, becomes smaller
than the width of the antenna equal to 2.2 μm. Figure 22d further characterizes
this wavelength conversion process. It shows the dependence of the spin-wave
wavelength at the output of the tapered part on the wavelength in the excitation
area. These data show that the discussed conversion process enables the reduction
of the wavelength by up to a factor of 15 and allows one to efficiently excite
spin waves with the wavelength below 1 μm, which is more than by a factor of 2
smaller than the antenna width. We also note that, as shown in [59], the conversion
process is not accompanied by noticeable additional losses associated with the spin-
wave reflection in the transition region, which makes it favorable for technical
applications.

Spin-Torque Transfer Effect and Spin Waves

Since the first demonstration [25, 27–29, 66, 67] of the possibility to excite
magnetization dynamics by spin-polarized electric currents due to the spin-transfer
torque (STT) effect [6, 7], dynamic spin-torque phenomena have become the subject
of intense research. The ability to control high-frequency magnetization dynamics
by dc currents is promising for the generation of microwave signals [68–72] and
propagating spin waves [16, 30, 73–75] in magnetic nanocircuits. Electronically
controlled local generation of spin waves is particularly important for the emerging
field of nanomagnonics, which utilizes propagating spin waves as the medium for
the transmission and processing of signals, logic operations, and pattern recognition
on nanoscale [3–5].
Initially, STT phenomena have been studied in 0D-nanodevices based on the
tunneling or giant magnetoresistance spin-valve structures, where STT is induced
by the electric current flowing through a multilayer that consists of a “fixed”
magnetic spin-polarizer and the active magnetic layer, separated by a nonmagnetic
316 S. O. Demokritov and A. N. Slavin

a) z 2 μm b)
y
cos (ϕ)
He 2.2 μm 1
L

on
0

ati
ag
0.5 μm

op
-1

Pr
c) 1
cos (ϕ)

-1
0 2 4 6 8 10
y, μm

d) 1.5
wavelength, μm
Converted

1.0

0 5 10 15 20 25
Excited wavelength, μm

Fig. 22 (a) Schematic of a tapered waveguide with the spin-wave antenna located in the wide
part. (b) Measured phase map of spin waves propagating in a 40 nm thick tapered waveguide
magnetized by the static field He = 900 Oe. The length of the tapered part L = 5 μm. Excitation
frequency is 9.8 GHz. (c) Section of the phase map along the waveguide axis. (d) Dependence of
the spin-wave wavelength at the output of the tapered part on the wavelength in the excitation area.
(© 2015 IEEE. Reprinted, with permission, from [47])

metallic or insulating spacer. In these structures, the electric charges must cross
the active magnetic layer to excite its magnetization dynamics. To enable current
flow through the active magnetic layers, STT devices operating with spin-polarized
electric current require that current-carrying electrodes are placed both on top and
on the bottom of the spin valve. To keep the electric current below a reasonable
limit, the devices should have sub-100-nm dimensions in both lateral directions.
An alternative approach to the implementation of STT devices that avoids
these shortcomings utilizes pure spin currents – flows of spin not accompanied
by directional transfer of electrical charge. This approach does not require the
flow of electrical current through the active magnetic layer, resulting in reduced
Joule heating and electromigration effects. One can also eliminate the electrical
leads attached to the magnetic layer to drain the electrical current, enabling novel
geometries and functionalities of the STT devices. Moreover, it becomes possible to
use insulating magnetic materials such as Yttrium Iron Garnet (YIG) [12, 76].
6 Spin Waves 317

Among the physical mechanisms for creation of pure spin currents, the spin-
Hall effect (SHE) [9–11, 77] plays the most important role so far. The effect is
generally significant in nonmagnetic materials with strong spin-orbit interaction,
such as Pt and Ta. An electrical current in these materials produces a spin current
in the direction perpendicular to the charge flow, due to a combination of spin-
orbit splitting of the band structure (intrinsic SHE), and the spin dependence of the
electron scattering on phonons and impurities (extrinsic SHE) [11, 77]. When a SHE
layer is brought in contact with a ferromagnetic film, the spin current flows through
the interface into the ferromagnet and exerts STT on its magnetization [78]. The
ability to exert STT on ferromagnets over extended areas is a significant benefit of
SHE. Indeed, when an in-plane current flows through an extended bilayer formed by
a SHE material and a magnetic film, the spin current produced by SHE is injected
over the entire area of the sample, which can be as large as several millimeters
[78]. This feature makes SHE uniquely suited for the control of the spatial decay
of propagating spin waves in 1D waveguides, when STT partially compensates the
natural magnetic damping.
The effect of pure spin current on the magnetization is similar to that of
spin-polarized electric currents. Both can be described by the Slonczewski-Landau-
Lifshitz-Gilbert equation [8]:

dM α dM β  
= −γ M × H eff + M× + 2 M × M × ŝ (25)
dt Ms dt Ms

where α is the Gilbert damping parameter, β is the strength of the spin-transfer


torque proportional to the spin current density, and ŝ is the unit vector in the
direction of the spin-current polarization. All other notations are similar to those
of Eqs. (1) and (2). In fact, Eq. (25) is an extension of Eq. (1), which takes into
account magnetic damping and the STT effect. Within this model, the third term
is mathematically very similar to the second one. Correspondingly, one expects
that spin current with an appropriate polarization can reduce magnetic damping of
propagating spin waves [78].
The above spin-wave damping compensation has been recently demonstrated
experimentally [79]. Figure 23 shows the schematic of the test devices. They
are based on a 20 nm thick YIG film grown by the pulsed laser deposition on
Gadolinium Gallium Garnet (GGG) (111) substrate. The film is covered by an
8 nm thick layer of Pt deposited using dc magnetron sputtering and the YIG/Pt
bilayer is patterned by e-beam lithography into a stripe waveguide with the width of
1 μm. The system is insulated by a 300 nm thick SiO2 layer, and a broadband
3 μm wide microwave antenna made of 250 nm thick Au is defined on top of
the system by the optical lithography. The waveguide is magnetized by the static
magnetic field He = 1000 Oe applied in its plane perpendicular to the long axis.
A dc electrical current I flowing in the plane of the Pt film is converted by the
SHE into the transverse spin accumulation (see inset in Fig. 23). The associated
pure spin current IS is injected into the YIG film resulting in a spin-transfer torque
on its magnetization. Depending on the relative orientation of the current and the
318 S. O. Demokritov and A. N. Slavin

I
Pt
Spin-wave Microwave IS
YIG
antenna current
M

He
Spin wave
I

z y Pt (8 nm) /
x YIG (20 nm)
waveguide
Probing
laser light GGG substrate

Fig. 23 Schematic of the experiment on compensation of the spin-wave damping by pure spin
current. Inset illustrates the generation of the pure spin current by the spin-Hall effect. (Reprinted
from [79], with the permission of AIP Publishing)

static magnetic field, the spin current either compensates or enhances the effective
magnetic damping in the YIG film.
The effects of spin current on the decay length of propagating spin waves were
performed by applying a microwave signal at the frequency corresponding, for the
given conditions, to a spin wave with the wavelength of about 5 μm, which can be
efficiently excited by the used 3 μm wide inductive antenna and possess sufficiently
large group velocity. The propagation of spin waves was mapped by rastering
the probing laser spot over the surface of the YIG waveguide with step sizes of
200 and 250 nm in the transverse and the longitudinal directions, respectively.
Figure 24a shows a representative map of the BLS intensity, proportional to the
local intensity in the spin wave, obtained for I = 2.55 mA. As seen from these data,
the spin wave propagates along the waveguide nearly uniformly without changing
its transverse profile (inset in Fig. 24a), which is a clear signature of the single-
mode propagation regime caused by the strong separation of the transverse modes
in a narrow waveguide. The intensity of the wave decreases by only 60% over the
propagation path of 10 μm. To characterize the decay length of spin waves and its
dependence on the current, we plot in Fig. 24b the dependences of the spin-wave
intensity on the propagation coordinate obtained for different dc currents in the Pt
layer. These data show that spin waves in the waveguide experience well-defined
exponential decay (note the logarithmic vertical scale) ∼ exp(−2y/ξ), where ξ is the
decay length defined as a distance over which the wave amplitude decreases by a
factor of e. As seen from the figure, the decay length strongly increases with the
increase of the dc current, as expected for the effect of spin current on the effective
magnetic damping.
6 Spin Waves 319

Fig. 24 (a) Normalized spatial intensity map of the propagating spin wave excited by the antenna.
The map was recorded for I = 2.55 mA. The mapping was performed by rastering the probing spot
over the area 1.6 by 10 μm, which is larger than the waveguide width of 1 μm. Dashed lines
show the edges of the waveguide. Inset shows the transverse profile of the spin-wave intensity. (b)
Dependences of the spin-wave intensity on the propagation coordinate for different currents, as
labeled, in the log-linear scale. Lines show the exponential fit of the experimental data. (c) Current
dependences of the decay length and the decay constant. Vertical dashed line marks IC . Solid line is
the linear fit of the experimental data at I < IC . The data were obtained at He = 1000 Oe. (Reprinted
from [79], with the permission of AIP Publishing)
320 S. O. Demokritov and A. N. Slavin

Figure 24c summarizes the results of the spatially resolved measurements. The
decay length (up-triangles) monotonously increases with the increase of I < IC
and then shows an abrupt decrease at I > IC in contradiction to naive expectations
that for large values of I the magnetic damping should be overcompensated by the
spin current, and the propagating spin wave should be amplified. This experimental
observation can be attributed to the strong nonlinear scattering of the propagating
spin waves from large-amplitude current-induced magnetic fluctuations, which have
been observed independently. To characterize the variation of the decay length
with current in detail, we plot in Fig. 24c its inverse value – the decay constant
(down-triangles), which is proportional to effective Gilbert damping constant αeff .
In agreement with the simple theoretical model assuming the linear variation of αeff
with current, the decay constant shows a linear dependence on I. By extrapolating
this dependence to I = 0, we obtain the propagation length at zero current
ξ0 = 2.4 μm. Additionally, one expects the linear dependence in Fig. 24c to cross
zero at I = IC , which corresponds to an infinitely large decay length under conditions
of the complete damping compensation. The data of Fig. 24c show, however, that
the linear fit yields the intercept value larger than IC . This disagreement can be
attributed to the Joule heating of the waveguide by the electric current in Pt resulting
in the significant reduction of the effective magnetization. Since the decay length is
proportional to the group velocity, which is known to decrease with the decrease
in Meff , the effects of the heating on the propagation length counteract those of the
spin current and do not allow one to achieve the decay-free propagation regime. We
note, however, that the maximum achieved propagation length of 22.5 μm is nearly
by a factor of two larger compared to the value of 12 μm estimated for a waveguide
made of a bare YIG film without Pt on top (α = 5 × 10−4 ).

Spin Waves in 0D

The STT effect discussed above is of a particular importance for magnetic systems
fully confined in all three directions. It is now well established that a spin-polarized
electric current or, alternatively, a pure spin current, injected into a ferromagnetic
layer through a nanocontact exerts a torque on the magnetization, leading to a
strongly localized microwave-frequency precession of magnetization, which can be
considered as a 0D spin-wave mode. This phenomenon can serve as a basis for the
development of tunable nanometer-size microwave oscillators, the so-called spin-
torque nano-oscillators (STNO) [25, 27–29, 66, 67]. The density of magnetic energy
in auto-oscillations excited by STT in a magnetic nanocontact could be very high.
Therefore, the STT-induced precession modes are, usually, strongly nonlinear. Also,
since the spin precession excited in a magnetic nanocontact is, usually, surrounded
by a 2D film or is coupled to a 1D waveguide, it may radiate propagating spin waves.
All these makes the phenomena connected with the STT-driven magnetization
dynamics multifarious and intriguing. This section is devoted to the 0D spin-wave
modes driven by spin-polarized electric or pure spin currents.
6 Spin Waves 321

Probing laser light

Top electrode
Current
flow
Insulator
nm
200

nm
500 Py film

Nanopillar
Bottom electrode

Fig. 25 Schematic of the studied STNO with an AFM image superimposed. The devices consist
of an extended 6 nm thick Permalloy free layer and an elliptical nanopillar formed by a 9 nm thick
Co70 Fe30 polarizing layer and a 3 nm thick Cu spacer. The nanopillar is located close to the edge of
the top electrode enabling optical access to the free layer for BLS microscopy. Magnetic precession
in the device is induced by dc current flowing from the polarizer to the free layer. The spatially
resolved detection of spin waves is accomplished by focusing the probing laser light into a 250 nm
spot, which is scanned over the surface of the Py film. (Reprinted from [31], with the permission
of Springer Nature)

Spin-Torque Nano-Oscillator (STNO) and Emitted Spin Waves

Let us consider an STNO shown in Fig. 25 [30]. The device is formed by a


nanocontact on an extended Permalloy (Py) film. The nanocontact is shaped as
an elliptical nanopillar formed by the nanopattered polarizing Co70 Fe30 layer and
a Cu spacer. A dc current I flowing from the polarizer to the Py film induces
local magnetization oscillations in this film. The nanocontact is located within 200
nanometers from the edge of the top device electrode, enabling optical access to the
Py film at larger distances. The spatially resolved detection of spin waves emitted by
STNO was performed by micro-focus BLS spectroscopy, as described above. The
probing laser light was focused onto the surface of the Py film and scanned in plane
to record two-dimensional maps of the spin-wave intensity.
The oscillation characteristics of STNOs were determined from the microwave
signals generated due to the magnetoresistance effect as shown in Fig. 26. The plots
of the power spectral density (PSD) illustrate the dependence of the oscillation
frequency on the bias current I for three different angles ϕ between the in-plane
bias magnetic field He = 900 Oe and the easy axis of the nanostructured polarizer.
The microwave generation starts at an onset current I = 2.5–3.5 mA that depends
on ϕ. The dependence of the generation frequency on current above the onset is
caused by the nonlinear frequency shift, due to a combination of the demagnetizing
322 S. O. Demokritov and A. N. Slavin

He 5° He He
25° 45°

8.2 8.2 8.2


Frequency, GHz
8.0 8.0 8.0

7.8 7.8 7.8

7.6 7.6 7.6

7.4 7.4 7.4

7.2 7.2 7.2


2 3 4 5 2 3 4 5 2 3 4 5
I, mA PSD,
-5 -4 -3 -2 -1
pW/MHz 10 10 10 10 10

Fig. 26 Pseudo-color logarithmic maps of the power spectral density (PSD) of the signal
generated by the device due to the magnetoresistance effect at different angles ϕ between the
in-plane magnetic field and the easy axis of the nanopillar, as labeled. (Reprinted from [31], with
the permission of Springer Nature)

effects in Py and the dipolar field of the structured Co70 Fe30 polarizer. For small ϕ,
the nonlinear shift is strongly negative. It becomes less pronounced with increasing
ϕ and changes to positive at small I and ϕ > 20◦ . The region of positive nonlinear
frequency shift is reduced at larger He and eventually disappears for He > 1200 Oe,
suggesting its origin from the dipolar field of the polarizer. The possibility to control
the nonlinear behaviors by varying the angle ϕ makes the studied STNOs uniquely
suited for the analysis of the effects of the nonlinearity on the spin-wave emission.
Figure 27 shows two-dimensional intensity maps of spin waves emitted by
STNO at I = 5 mA, measured for different in-plane directions of the applied field
He = 900 Oe. As seen in Fig. 27, the emission mainly occurs in the direction
perpendicular to the in-plane field, regardless of its orientation, the generation
frequency, or the magnitude of the nonlinear frequency shift. We note that although
the sign of the nonlinear shift is expected to be important for the efficiency of
spin-wave emission, the maps of Fig. 27 corresponding to significantly different
nonlinear behaviors of the STNO (see Fig. 26) differ predominantly by the direction
of emission, which rotates together with the field.
Figure 28 illustrates the spin-wave characteristics determined at the location of
the maximum spin wave intensity. Figure 28a–c show the BLS spectra of the emitted
spin waves for I increasing from 3 to 5 mA, together with the spectrum of the
thermal spin waves. The spectra exhibit a small nonlinear frequency shift at ϕ = 45◦ ,
which increases as ϕ is reduced, in agreement with the electrical measurements
shown in Fig. 26. Figure 28d summarizes the dependences of the frequency-
integrated spin-wave intensity on the current I. As seen from these data, the intensity
of the emitted spin waves increases linearly with current for the angle ϕ = 45◦
characterized by a small nonlinear frequency shift. In contrast, the data for ϕ = 25◦
6 Spin Waves 323

a b He
He

He
He

500 nm

c d
He

He He

He

Normalized
intensity 0.0 0.2 0.4 0.6 0.8 1.0

Fig. 27 Normalized color-coded intensity maps of spin waves emitted by the STNO, recorded
at different angles ϕ between the in-plane magnetic field He = 900 Oe and the easy axis of the
elliptical nanopillar: (a) ϕ = 5◦ o, (b) ϕ = 25◦ o, (c) ϕ = 45◦ o, (d) ϕ = − 45◦ o. The bias current is
I = 5 mA. The schematic of the top electrode is superimposed on each map, with a cross indicating
the location of the nanocontact. The intensity maps acquired at I = 0 were subtracted to eliminate
the contribution from the thermal spin waves. Arrows show the direction of the static magnetic
field, and the dashed lines indicate the direction of the spin-wave emission. (Reprinted from [31],
with the permission of Springer Nature)

and ϕ = 5◦ exhibit a decrease of the spin-wave intensity starting from a certain


value of current that decreases with decreasing ϕ. These findings are correlated
with a larger nonlinear frequency shift, resulting in more significant reduction of the
emission frequency far below FMR. In contrast, magnetoresistance measurements
(Fig. 26) showed similar monotonic increases of generated power for all three
configurations. Therefore, the decrease in the BLS intensity is associated with a
decreased emission efficiency rather than a reduced amplitude of the oscillation in
the nano-contact area.
However, one should admit that decay length of the emitted waves was rather
small, below 500 nm. Further studies [16, 73, 80] have shown that the spin waves
emitted in these experiments have an evanescence nature, since their frequency were
slightly below the spin-wave spectrum of the surrounding Py film. As demonstrated
in Fig. 29, microwave parametric pumping can be used as a mechanism for the
transfer of the generated microwave energy into the desirable spectral range above
the FMR frequency [80]. This effect enables an increase of the propagation length of
324 S. O. Demokritov and A. N. Slavin

a He 5° b He

BLS intensity, a.u.

BLS intensity, a.u.


25°

7 8 9 10 11 7 8 9 10 11
Frequency, GHz Frequency, GHz

c He d
45°
BLS intensity, a.u.

Integral intensity

7 8 9 10 11 3 4 5
Frequency, GHz I, mA

Fig. 28 (a–c), Dependence of BLS spectra on the current for different in-plane directions of
He = 900 Oe, as indicated. Shadowed regions show the spectrum of the thermally excited spin
waves determined by measurement at I = 0. Color lines show the spectra acquired at the currents
of 3.0 mA (black), 3.5 mA (blue), 4.0 mA (green), 4.5 mA (red), and 5.0 mA (pink). Dashed
vertical lines mark the frequency of the ferromagnetic resonance (FMR). (d) Dependences of the
integrated intensity of emitted spin waves on current for ϕ = 5◦ (triangles), 25◦ (squares), and 45◦
(dots). (Reprinted from [31], with the permission of Springer Nature)

spin waves emitted by STNOs: the decay length of 540 nm for the auto-emission was
increased to 940 nm for the pumping-induced emission. Moreover, the phenomenon
of the pumping-induced emission does not disturb the unique directionality found
for the emission in the auto-oscillation regime, as illustrated by Fig. 29.

Spin-Hall Nano-Oscillator (SHNO)

In the previous section, a STNO driven by spin-polarized electric current was


considered. Another possibility to inject angular momentum into a magnetic system
is utilization of pure spin current. As it has been already mentioned above, the
application of pure spin current has numerous advantages compared to the spin-
polarized electric current when the excitation of a large-amplitude 0D spin-wave
modes is discussed. A complete compensation of damping by the spin current
6 Spin Waves 325

Fig. 29 Pseudocolor spatial intensity maps of the emitted spin waves, acquired at I = 5 mA.
A schematic of the top electrode and a cross indicating the location of the nanocontact is
superimposed on each map. (a) Spin-wave auto-emission, in the absence of the external pumping
microwaves. (b) Spin-wave emission under influence of parametric pumping. Note an extended
spin-wave propagation area for (b). (Reprinted with permission from [80], © 2011 by the American
Physical Society)

appears to be a straightforward extension of the damping reduction, described in


Sect. “Spin-Torque Transfer Effect and Spin Waves”. However, as the compensation
point is approached, additional nonlinear damping emerges due to the nonlinear
interactions among different dynamical modes enhanced simultaneously by the spin
current, preventing the onset of auto-oscillation. Since magnon-magnon scattering
rates are proportional to the populations of the corresponding modes, detrimental
effects of nonlinear damping can be avoided by selectively suppressing all the
modes, except for the ones that can be expected to auto-oscillate. To achieve
selective suppression, the frequency-dependent damping caused by the spin-wave
radiation was used. To take advantage of this radiative damping, the spin current
was locally injected into an extended magnetic film, in contrast with the geometry
described in Sect. “Spin-Torque Transfer Effect and Spin Waves”. In fact, the local
spin current enhances a large number of dynamical modes, but those having higher
frequencies, and, consequently, higher group velocities, quickly escape from the
active region, which results in their efficient suppression by the radiation losses.
Meanwhile, the modes at frequencies close to the bottom of the spin-wave spectrum
have a much smaller group velocity, and, therefore, minimum radiation losses.
The scheme of our experiment with pure spin current is shown in Fig. 30a [31].
The studied device is formed by a bilayer of a 8 nm thick film of Pt and a 5 nm thick
film of Py patterned into a disk with a diameter of 4 μm. Two 150 nm thick Au
electrodes with sharp points separated by a 100 nm wide gap are placed on top of
the bilayer, forming an in-plane point contact. The sheet resistance of the bilayer is
nearly two orders of magnitude larger than that of the Au electrodes. Consequently,
the electrical current induced by voltage between the electrodes should be strongly
localized in the gap. Indeed, a calculation of the current distribution through a
326 S. O. Demokritov and A. N. Slavin

Fig. 30 (a) a
Scanning-electron
DC
microscopy image of the test Probing laser Current
spin-Hall nano-oscillator. The light
device consists of a 4 μm
diameter disk formed by a
8 nm thick Pt on the bottom
and a 5 nm thick Py layer on
top, covered by two pointed
Au(150 nm) electrodes Py(5)Pt(8)
He Disk
separated by a 100 nm gap.
(b) Normalized calculated Au(150)
distribution of current 1μ Top electrodes
m
through the section of the b
device shown in the inset by a
dashed line. (Reprinted from 1
current density
Normalized

[30], with the permission of


Springer Nature) 0.5 z 250 nm

0
-2 -1 0 1 2
z, μm

section across the middle of the gap (Fig. 30b) shows that most of the current flows
through a 250 nm wide Pt strip. This electric current creates a pure spin current
flowing into Py, due to the spin-Hall effects. The spin current injected into Py exerts
spin-transfer torque on its magnetization. As a result, the damping is compensated,
and the dynamic magnetic modes are enhanced.
Figure 31 shows the BLS spectra obtained with the probing spot positioned in
the center of the gap between the electrodes, at different values of the dc current
I. At I = 0, the BLS spectrum exhibits a broad peak corresponding to incoherent
thermal magnetization fluctuations in the Py film (Fig. 31a). As this thermal peak
grows with increasing current, its rising front becomes increasingly sharper than
the trailing front, consistent with the preferential enhancement of the low-frequency
modes. Analysis of the dependence of the frequency-integrated BLS intensity on
current (Fig. 31b) shows that the intensity of magnetic fluctuations diverges as the
current approaches a critical value of Ic ≈ 16.1 mA. In contrast to confined systems
driven by spatially uniform spin currents [81], the intensity of fluctuations does
not saturate as the current approaches Ic , indicating that the nonlinear processes
preventing the onset of auto-oscillations are avoided.
At I ≥ Ic , a new peak appears in the BLS spectrum below the thermal peak, as
indicated in Fig. 31a by an arrow. The calculated current density in the center of the
gap at the onset is 3 × 108 A/cm2 , which is only slightly larger than the extrapolated
value 1 × 108 A/cm2 obtained for a similar system without radiation losses [81].
Since this peak is not present in the thermal fluctuation spectrum, we can conclude
that it corresponds to a new auto-oscillation mode that does not exist at I < Ic . The
peak rapidly grows and then saturates above 16.3 mA (Fig. 31c–d). Comparing the
6 Spin Waves

Fig. 31 (a) BLS spectra of thermal fluctuation amplified by the spin current at currents below the onset of auto-oscillation. (b) Integral intensity of amplified
thermal fluctuations and its inverse versus current. Both dependencies are normalized by their values at I = 0. (c) BLS pectra of the magnetization auto-
oscillation driven by the spin current. Filled areas are the results of fitting by the Gaussian function. Note, that the spectral widths are determined by the
327

resolution of the BLS setup. (d) The intensity and the center frequency of the auto-oscillation peak versus current. Curves are guides for the eye. Reprinted
from [30], with the permission of Springer Nature
328 S. O. Demokritov and A. N. Slavin

spectra for I = 16.1 mA and 16.3 mA, we see that the onset of auto-oscillations is
accompanied by a decrease in the intensity of thermal fluctuations, suggesting that
the energy of the spin current is mainly channelled into the auto-oscillation mode.
The spectral width of the auto-oscillation peak characterizing the coherence
of auto-oscillations decreases just above Ic , and stabilizes above 16.3 mA. Note
that the linewidth in the spectra shown in Fig. 31a and c is determined by the
spectral resolution of our optical technique under usual conditions. Additional
measurements at our instrument’s ultimate spectral resolution of 60 MHz show that
the actual linewidth in the saturated regime is below this value, suggesting a high
degree of coherence of the observed auto-oscillation mode.
The frequency of the auto-oscillation peak monotonically decreases with increas-
ing I (Fig. 31d). We note that the generated frequency is significantly below the
frequencies of magnetic fluctuations even at the onset of auto-oscillations. We draw
three important conclusions based on this observation. First, the auto-oscillation
mode does not belong to the thermal spin-wave spectrum. Second, this mode is
formed abruptly at the onset current, and not by gradual reduction of frequency
from the spin-wave spectrum due to the red nonlinear frequency shift. Third, since
the energy can be radiated only by propagating spin waves and there are no available
spin-wave spectral states at the auto-oscillation frequency, the auto-oscillation mode
is not influenced by the radiation losses.
To determine the spatial profile of the auto-oscillation mode, we performed
two-dimensional mapping of the dynamic magnetization at the frequency of auto-
oscillations, by rastering the probing laser spot in the two lateral directions and
simultaneously recording the BLS intensity. An example of the obtained maps
is presented in Fig. 32. These data show that the auto-oscillations are localized
in a very small area in the gap between the electrodes. The spatial distribution

Fig. 32 Normalized
color-coded map of the
measured BLS intensity over
the auto-oscillation area, and
two orthogonal sections
through its center. Symbols
are the experimental data, and
filled areas under solid curves
are the results of fitting by a
Gaussian function. Dashed
lines on the map show the
contours of the top electrodes.
The data were recorded at
I = 16.2 mA. (Reprinted from
[30], with the permission of
Springer Nature)
6 Spin Waves 329

of the BLS intensity is well described by a Gaussian function with the width
of 250 ± 10 nm, close to the diameter of the probing laser spot. The measured
spatial distribution is a convolution of the actual spatial profile with the instrumental
function determined by the shape of the laser spot. Therefore, we estimate that
the size of the auto-oscillation region is less than 100 nm, significantly smaller
than the characteristic size of the current localization (Fig. 30b). Therefore, we
conclude that the auto-oscillation area is determined not by the spatial localization
of the driving current, but by the nonlinear self-localization processes defining the
geometry of a standing spin-wave “bullet” [82]. We emphasize that the observed
quick saturation of the intensity of the auto-oscillation peak above the onset and
its monotonic red frequency shift are the intrinsic characteristics of the “bullet”
mode. Only one “bullet” mode exists at the frequency of the auto-oscillations and
this frequency is well separated from the continuous spectrum of non-localized spin
waves, Therefore, our findings provide strong evidence for that auto-oscillations
involve only a single mode in the studied system.

Nature of Spin Wave Modes Excited in 0D Magnetic Nanocontacts

The nature of the auto-oscillation spin wave mode excited by either spin-polarized
or pure spin current in magnetic nanocontacts (0D objects) is of a fundamental
importance for the current-induced magnetization dynamics.
The first theoretical analysis of the nature of the spin-wave eigen-mode excited
by spin-polarized current in a nano-contact geometry was performed by J. Slon-
czewski [8]. He developed a spatially nonuniform linear theory of spin wave
excitations in a nano-contact, where the “free” ferromagnetic layer is infinite in
plane, while the spin-polarized current traversing this layer has a finite cross-section
S = π Rc2 , where Rc is the contact radius. Considering a perpendicularly magnetized
nano-contact Slonczewski showed that in the linear case the lowest threshold
of excitation by spin-polarized current is achieved for an exchange-dominated
propagating cylindrical spin wave mode having wave number q0 = 1.2/Rc and
frequency [8]:

ω (q0 ) = ω0 + Dex q02 . (26)

Here ω0 is the ferromagnetic resonance (FMR) frequency in the magnetic film,


Dex = ωM lex 2 , ω ≡ 4π γ M , γ is the gyromagnetic ratio for electron spin, l
M S ex =
 2
1/2
A/2π MS is the exchange length, A is the exchange constant, and MS is the
value of the saturation magnetization.
It was also shown that the threshold current Ith in such a geometry consists of
two additive terms: the first one arises from the radiative loss of energy carried by
the propagating spin wave out of the region of current localization, while the second
one is caused by the usual energy dissipation in the current-carrying region:
330 S. O. Demokritov and A. N. Slavin

D (H )
lin
Ith = 1.86 + . (27)
σ Rc2 σ

Here σ = εgμB /2eMS dS whereε is the spin-polarization efficiency defined in


[8], g is the spectroscopic Lande factor, μB is the Bohr magneton, e is the modulus
of the electron charge, d is the thickness of the “free” magnetic layer, S is the cross-
section area of the nano-contact), and (H) is the spin wave damping dependent
on the bias magnetic field H. It turns out that for a typical nano-contact of the
radius Rc = 20 − 30 nm the radiative losses are about one order of magnitude
larger than the direct energy dissipation, and should give the main contribution into
the threshold current. This result, however, contradicts experimental observations
(see, e.g., [83]): the experimentally measured magnitude of the threshold current
in an in-plane magnetized nano-contact is much smaller than the value predicted
by Eq. (27), although the dependence of this current on the magnetic field H is
satisfactory described by this equation.
In this section we present a spatially nonuniform nonlinear theory of spin wave
excitation by spin-polarized current in a nano-contact geometry for the case of the
in-plane magnetization [82]. We show that in an in-plane magnetized magnetic film
the competition between the nonlinearity and exchange-related dispersion leads to
the formation of a stationary two-dimensional self-localized nonpropagating spin
wave mode. Such nonlinear self-localized wave modes in two- or three-dimensional
cases are conventionally called wave “bullets” [84]. The frequency of this spin wave
“bullet” is shifted by the nonlinearity below the spectrum of linear spin waves
and, therefore, this nonlinear mode has an evanescent character with vanishing
radiative losses, which leads to a substantial decrease of its threshold current Ith
in comparison to the linear propagating mode shown in Eq. (27).
To describe the generation of a spin wave bullet by the spin-polarized current
we consider a “free” ferromagnetic layer, infinite in y − z plane and having finite
thickness d in the x direction (d is assumed to be sufficiently small for us to consider
that the magnetization M is constant along the film thickness, and that the dipole-
dipole interaction can be described by a simple demagnetization field). We assume
that the internal magnetic field H = Happ + Hex , consisting of the applied Happ and
interlayer exchange Hex fields, is applied in the z direction in the film plane. Using
the standard Hamiltonian spin-wave formalism [33], which has been successfully
used to develop a spatially uniform nonlinear model of spin wave generation by
spin-polarized current [85, 86], one can derive an approximate equation for the
dimensionless complex spin wave amplitude b ≡ b(t, r):

∂b  
= −i ω0 b − DD
b + N |b|2 b − b + f (r/Rc ) σ I b − f (r/Rc ) σ I |b|2 b.
∂t
(28)

Here ω0 ≡ ωH (ωH + ωM ) is the linear FMR frequency, (ωH ≡ γ H,
DD ≡ (2A/MS ) ∂ω0 /∂H = (2γ A/MS )(ωH + ωM /2)/ω0 is the dispersion coefficient
6 Spin Waves 331

for spin waves,


is the two-dimensional Laplace operator in the film plane,
N = − ωH ωM (ωH + ωM /4)/ω0 (ωH + ωM /2) is the coefficient describing nonlinear
frequency shift, and  ≡ α G (ωH + ωM /2) is the spin wave damping rate (α G is
the dimensionless Gilbert damping parameter). The dimensionless function f (x)
describes the spatial distribution of the spin-polarized current. The dimensionless
spin wave amplitude b is connected with the z-component of the magnetization by
the equation |b|2 = (MS − Mz )/2MS .
Equation (28) differs from the Eq. (9) in [8] (which resulted in the solution
(27)) by the presence of two additional nonlinear terms: the term containing the
coefficient N and describing a nonlinear frequency shift of the excited mode, and
the last term describing the current-induced positive nonlinear damping that stops
the increase of the amplitude of the excited mode at relatively large currents. Also,
since the Eq. (28) was obtained as a Taylor expansion it is exactly correct only for
sufficiently small spin wave amplitudes |b| < 1.
Without damping and current terms ( = 0, I = 0) Eq. (28) coincides with
the well-known (2 + 1)-dimensional nonlinear Schrödinger equation (NSE) [87].
In the considered case of an in-plane magnetized film the nonlinear coefficient N
is negative, and the nonlinearity and dispersion satisfy the well-known Lighthill
criterion ND < 0 (i.e., they act in opposite directions), and the NSE has a nonlinear
self-localized radially symmetric standing solitonic solution (or the solution in the
form of a standing spin wave bullet)

b (t, r) = B0 ψ (r/) e−iωt , (29)

where dimensionless function ψ(x), having maximum value of 2.2 at x = 0,


describes the profile of the bullet. This function is the localized solution of the
equation

1
ψ + ψ + ψ 3 − ψ = 0, (30)
x

which has to be found numerically (see e.g., [84]).


In Eq. (29) B0 , , and ω are the characteristic amplitude, characteristic size, and
frequency of the bullet, respectively. Among these three parameters only one is
independent. Taking the amplitude B0 as an independent parameter, we can express
the two other parameters as

|D/N|
ω= ω0 + NB 20 , = . (31)
B0

We would like to stress that the frequency of the spin wave bullet lies below
the linear frequency ω0 of the ferromagnetic resonance (see Eq. (31), and note that
N < 0), i.e., outside the spectrum of linear spin waves. This is the main reason for
the self-localization of the spin wave bullet, as the effective wave number of the spin
332 S. O. Demokritov and A. N. Slavin

wave mode with frequency (6) is purely imaginary. It also follows from Eq. (29) and
the expansion condition |b| < 1 that the maximum magnitude of B0 for which our
perturbative approach is still correct is B0 = 0.46.
It is well known [87] that the bullet-like solutions of (2 + 1)-dimensional NSE
are unstable with respect to the small perturbations: the wave packets having the
bullet shape Eq. (29), but amplitudes smaller than B0 , decay due to the dispersion
spreading, while the wave packets having amplitudes higher than B0 collapse due
to the nonlinearity. At the same time, Eq. (28) with both Gilbert dissipation  and
current I is a two-dimensional analog of a Ginzburg-Landau equation that is known
to have stable localized solutions (see, e.g., review [88]). One can assume that for
a small damping rate  and current I the full nonconservative eq. (28) will have a
bullet-like solution, only slightly different from the exact solution Eq. (29) of the
conservative NSE equation. It is clear, however, that not all of such solutions can be
supported in our case. For example, small-amplitude bullets, for which  > > Rc ,
practically do not interact with the spatially localized current and will decay due to
the linear dissipation. The large-amplitude (B0 ≥ 1) bullets, on the other hand, will
also decay because the effective damping  − σ I(1 − |b|2 ) for them changes sign
and becomes positive.
The excitation threshold of the spin wave bullet mode was calculated in [82] and
the minimum value of this threshold turned out to be equal to the second term in Eq.
(27), i.e., sobstantially lower than the threshold of excitation of the propagating
spin-wave mode in the perpendicularly magnetized magnetic nanocontact
Eq. (27).
To find the spatial profile of the spin-wave bullet mode Eq. (28) was solved
numerically. The results of comparison of the spin-wave excitation profiles at
the threshold obtained for a typical set of experimental parameters [83] from the
analytical solution Eq. (29) (solid black line) and numerical solution of Eq. (28)
(black dots) are shown in Fig. 33. One clearly sees that the numerical profile of
the nonlinear eigen-mode is practically indistinguishable from the approximate
“bullet-like” profile, so the “bullet” model works exceptionally well in this case.
For comparison we present in Fig. 33 the spatial profile of the Slonczhewski-like [8]
linear mode, that is obtained from the solution of Eq. (28) where the nonlinear terms
(terms containing |b|2 ) are omitted (red line). The amplitude of this linear mode at
the threshold is vanishingly small, |b(r)|2 → 0. We also present the normalized
spatial profile of a bullet mode above the thershold numerically calculated form Eq.
(28), to show that with the increase of the bullet amplitude its width decreases in
accordance with the experssion shown in Eq. (31) (blue line).
As it was mentioned above, the bullet mode is excited in an in-plane magetized
nanocontact, while the linear propagating spin wave mode (Slonczewski mode [8])
could be excited in a perpendicularly magnetized nanocotact. It is also well-known
that in the case when the direction of the external bias magnetic field varies from in-
plane to the perpendicular the coefficient of the nonlinear frequency shift N changes
its sign from negative to positive [89, 90]. Therefore, it is interesting to see the nature
of the spin-wave mode excited by a spin-polarized (or pure spin) current under the
oblique magnetization of a nanocontact.
6 Spin Waves 333

1.0

Normalized power |b(r)|2/|b(0)|2


Linear mode

0.8
Nonlinear "bullet"
0.6
Nonlinear "bullet"
0.4 above the threshold

0.2

0.0
0.0 0.5 1.0 1.5 2.0 2.5 3.0
Normalized distance r/Rc

Fig. 33 Normalized profiles of the spin wave mode generated by spin-polarized current at (the
threshold: solid black line – bullet profile (29), circles – result of the numerical solution of Eq.
(28), red line – profile of the linear eigen-mode calculated from the linearized Eq. (28). Blue line
demonstrates the numerically calculated profile of the bullet mode above the thershold. Vertical
dash-dotted line shows the region of the current localization. The parameters are: 4π M0 = 16.6 kG,
Happ = Hex = 5 kOe, A = 2.85 · 10−6 erg/cm, α G = 0.015, d = 1.2 nm, Rc = 25 nm, ε = 0.3

The analytic [91], numerical [92], and experimental [93] studies of the spin wave
excitation in current-driven magnetic nanocontacts were successfully performed,
and confirmed the conclusions of the above-presented theory.
In particular, the experimental study [93] was performed on a nanocontact
2Rc = 40 nm to the thin film tri-layer Co81 Fe19 (20 nm)/Cu(6 nm)/Ni80 Fe20 (4.5 nm),
patterned into a 8 μ m × 26μ m mesa. On top of this mesa, a circular Al nanocontact
was defined through SiO2 using e-beam lithography and an external magnetic field
of a constant magnitude (μ0 He = 1.1 T) was applied to the sample at an angle
θ e with respect to the film plane. Microwave excitations were only observed for a
single current polarity, corresponding to electrons flowing from the “free” (thinner)
to the “fixed” (thicker) magnetic layer. All measurements were performed at room
temperature.
Figure 34 shows the detailed angular dependence of the microwave frequencies
generated at a constant current of I = 14 mA and a constant magnetic field amplitude
of μ0 He = 1.1 T. The generated frequencies are approximately independent of
the magnetic field angle up to about θ e = 35◦ , and then decrease from about
35 GHz to 10 GHz with the increasing angle. The most important feature of the
results presented in Fig. 34 is the existence of two distinct and different modes for
sufficiently small values of θ e – linear propagating spin wave Slonczewski mode,
having a higher frequency, and a nonlinear self-localized spin-wave bullet mode
having a lower frequency. The frequencies of these two modes differ by about
2.5 GHz at angles up to θ e = 40◦ , and then they start to approach each other
up to θ e ≈ 55◦ where the mode having lower frequency completely disappears.
334 S. O. Demokritov and A. N. Slavin

40
Propagating mode
Spin wave bullet

30
f [GHz]
40

20 f [GHz] 30

20

10 10

0
0 15 30 45 60 75 90
Applied field agle θe [deg]
0
0 15 30 45 60 75 90
Applied field agle θe [deg]

Fig. 34 Experimental frequencies of the current-induced spin wave modes as a function of the
applied field angle θ e between the external bias magnetic field and the nanocontavt plane at I = 14
mA and μ0 He = 1.1 T (symols). Inset: theoretically calculated frequencies of the propagating
(upper curve) and “bullet” (lower curve) modes at the threshold of their excitation for nominal
parameters of the nanocontact sample. (Reprinted with permission from [82], © 2005 by the
American Physical Society)

The behavior of the frequencies of the current-induced spin wave modes shown in
Fig. 34 remains the same for any magnitude of the bias current that is larger than
I ≈ 10 mA. This behavior is also qualitatively similar to the behavior of the mode
frequencies derived from analytic theory (see Fig. 2 in [91] and the inset in Fig. 34)
and from the numerical simulation (see Fig. 4 in [92]).
The experimental threshold currents for the two excited spin wave modes shown
in Fig. 35 as functions of θ e were determined using the method proposed in [90]. The
graph in Fig. 35 only shows the threshold currents determined from experiment for
the magnetization angles 20◦ < θ e < 80◦ , since outside this range the signal was too
weak to allow reliable determination of the excitation threshold. It is clear that the
lower-frequency (bullet) mode has a lower threshold current at low magnetization
angles. As the angle increases, the threshold currents for the two modes gradually
approach each other and become essentially equal close to the critical angle
θ e ≈ 50◦ , where the low-frequency mode disappears. These experimental data are
also qualitatively similar to the threshold curves calculated analytically in [91] (see
solid lines in Fig. 35) and the similar curves simulated numerically (see Fig. 3 in
[92]). The inset in Fig. 35 shows the numerically calculated profile of the both
modes: nonlinear self-localized bullet mode (left frame) and linear propagating
mode (right frame). Thus, the results of the laboratory experiment [93] and the
results of the numerical simulations [92] fully confirmed the theoretical ideas [8,
82] about the possibility of current-induced excitation of two qualitatively different
spin wave modes in magnetic nanocontacs.
6 Spin Waves 335

30

25 100

Threshold current lth [mA]


100
0 0
y [nm] x [nm]
200 200
20 −100 −100 0 0 x [nm]
y [nm]
−200 −200

15

10

Propagating mode
5
Spin wave bullet

0
0 15 30 45 60 75 90
Applied field agle qe [deg]

Fig. 35 Measured threshold current for the propagating (empty triangles) and bullet (filled circles)
modes as a function of the applied field angle θ e . Solid lines: thresholds for the same modes versus
applied field angle theoretically calculated using the formalism of [91] (red line – bullet mode,
blue line- propagating mode). Upper inset demonstrates the numerically calculated spatial profiles
of the bullet mode (left) and propagating mode (right). (Reprinted with permission from [82], ©
2005 by the American Physical Society)

It is important to note that the excitation of self-localized evanescent spin wave


bullet modes by pure spin current was subsequently observed in several independent
experiments [31, 94, 95].
Another example of an interesting and unusual solitonic spin wave mode that
can be excited by spin-polarized current is given by the so-called spin-wave droplet
(or spin-wave droplet soliton) existing in perpendicularly magnetized nanocontacts
having a large perpendicular magnetic anisotropy (PMA) [96–98]. As it was first
demonstrated analytically in [99, 100], the Landau-Lifshitz equation for magnetic
films can sustain a family of so-called magnon drop solitons, provided there is no
spin wave damping.
While any realistic magnetic system always exhibits some spin wave damping,
hence making magnon drops unrealistic, it was demonstrated theoretically in [96]
that in a current-driven magnetic nanocontact with PMA, where the spin wave
damping is completely compensated by the STT effect [8], it would be possible
to excite a magnon-drop-like excitations. In contrast to the conservative magnon
drops, these so-called magnetic droplets are strongly dissipative, relying not only
on the zero balance between the exchange and anisotropy, but also on the balance
between the negative damping created by the STT effect and the positive nonlinear
damping in the current-driven magnetic material. As a consequence, out of a large
family of magnon drops, the additional net zero damping condition singles out a
particular magnetic droplet with both a well-defined frequency and a well-defined
direction of the dynamic magnetization at the center of the excited droplet. Note
336 S. O. Demokritov and A. N. Slavin

that at large droplet amplitudes the direction of magnetization at the droplet center
could be almost completely opposite to the similar direction at the droplet periphery
(see Fig. 2 in [96]). These rather exotic nonlinear dynamical magnetic modes
were observed experimentally in [97], and a more detailed description of magnetic
droplets is presented in [98].

Coupling of a STNO and 1D Spin-Wave Waveguide to Each Other

In the previous sections, we have demonstrated that STNO devices, employing 0D


spin-wave modes can convert the energy of direct electrical current into propagating
spin waves. We have also noticed that it is difficult to achieve frequency matching
of STNO with the propagating spin waves, since the large-amplitude spin wave
modes in STNOs are frequency shifted due to nonlinear properties of spin waves
with respect to characteristic frequencies of 1D and 2D propagating spin waves.
However, if one uses a spin-wave waveguide of a particular geometry as described
below, efficient matching between such waveguides and STNOs can be achieved.
This matching is realized by taking advantage of the dipolar magnetic field within
the waveguide, which acts on 1D propagating spin-wave modes [16].
Figure 36a shows the layout of the studied device. A point-contact STNO
is comprised of a multilayer Cu(4)/Co70 Fe30 (4)/Au(150) shaped into an elliptic
nanopillar with dimensions of 120 nm × 40 nm fabricated on top of an extended
5 nm thick Permalloy (Py) film. Additionally, the device incorporates a 5 nm thick
and 200 nm wide Co70 Fe30 nanostripe below the Py film. The device is magnetized
by a static magnetic field He = 800 – 1200 Oe applied in the plane of the Py film
perpendicular to the CoFe nanostripe.
Figure 36b shows the characteristics of the oscillation of STNO determined
by the standard electronic spectroscopy measurements. Above the onset current
of about 3.5 mA, both the amplitude and the frequency of the auto-oscillations
exhibit a smooth dependence on current, indicating a single-mode operation of
the STNO. Correspondingly, Fig. 36c shows representative BLS spectra recorded
with the probing laser spot positioned above the CoFe nanostripe. Note here
that by comparing Fig. 36b and c, one can conclude that the frequency of
the microwave signal is twice the frequency of the magnetization oscillation
measured by BLS, since the former is due to the quadratic magnetoresistance
effect.
While the BLS spectra acquired above the CoFe nanostripe clearly show the
signals resulting from the STNO oscillation, no such signals were detected away
from the nanostripe. This observation indicates that the STNO can efficiently
generate 1D spin waves, propagating along the CoFe nanostripe, but a radiation
of 2D spin waves into the free Py film is inefficient.
To understand this phenomenon, one has to consider the effects of the dipolar
field of the CoFe nanostripe on the internal field in the magnetic layers. Both
micromagnetic simulations and studies of spin-wave spectra of thermal fluctuations
6 Spin Waves 337

a Top
electrode Current

200 nm

Py (5) film
z
CoFe (5) Nano-
nano-wire y pillar
He
Bottom Cu(40)
electrode

b c
2.5
7 mA 1.0 7 mA
2.0
BLS intensity, a.u.
PSD, pW/MHz

0.8
1.5
0.6 5.5 mA
6 mA
1.0
0.4
5 mA 4 mA
0.5 0.2
4 mA
0 0.0
13 14 15
Frequency, GHz Frequency, GHz

Fig. 36 (a) Layout of the studied STNO with an incorporated waveguide. Inset: SEM micrograph
of the device. He is the static magnetic field. Numbers in parentheses indicate the thicknesses of
the layers in nanometres. (b) Spectra of the current-induced oscillations of the STNO measured
by a spectrum analyzer at different driving dc currents, as indicated. (c) BLS spectra recorded at
different driving currents measured by positioning the probing laser spot on the nano-waveguide.
Note that the spectral widths are determined by the resolution of the BLS setup. (Reprinted from
[16], with the permission of Springer Nature)

show that the internal field is significantly reduced in the magnetic film in the region
of the CoFe nanostripe, as compared to the magnitude away from the nanostripe.
The reduction of the internal field results in lowering of the local spin-wave
spectrum, creating a one-dimensional channel with allowed spin-wave frequencies
below the bottom of the spectrum in the free Py film. Low-frequency magnons
excited by STNO are directionally guided along the CoFe nanostripe, since there
are no states available at these frequencies in the free Py film. Thus, the static-
field channel induced by the CoFe nanostripe plays the role of a compound dipolar
spin-wave waveguide formed by the strongly exchange-coupled bilayer of the CoFe
nanostripe and the Py film on top of it.
338 S. O. Demokritov and A. N. Slavin

Fig. 37 (a) Normalized decay-compensated spatial map of the spin-wave intensity. The positions
of the top device electrode and the CoFe nanostripe are schematically shown. (b) Measured
dependence of the integral spin-wave intensity on the propagation coordinate (symbols), on
the log-linear scale. The line shows the result of the fitting of the experimental data by the
exponential function. (c) Distribution of the spin-wave intensity in the section transverse to the
nano-waveguide. Symbols are experimental data, curve is a fit by the Gaussian function. w is
the full width at half maximum of the transverse intensity profile. (d) Dependence of w on
the propagation coordinate. Symbols are experimental data, horizontal line is the mean value.
(Reprinted from [16], with the permission of Springer Nature)
6 Spin Waves 339

The measured propagation characteristics of spin waves in the nano-waveguide


are illustrated in Fig. 37. Figure 37a shows the normalized spatial map of the
BLS intensity, which is proportional to the local spin wave intensity. The map was
recorded at a constant dc current of 5 mA by rastering the probing laser spot over
a 1.6 μm by 1.6 μm area with the step size of 100 nm. To highlight the transverse
profile of the propagating wave, the spatial decay in the direction of propagation was
compensated by normalizing the signal with the integral over the transverse section
of the map (along the z-coordinate). The map of Fig. 37a clearly shows that the spin
wave energy is concentrated entirely in the nano-waveguide, i.e., spin waves are
guided by the field-induced channel without noticeable losses associated with the
radiation of energy into the surrounding free Py film.
The BLS intensity integrated over the transverse section of the map exhibits a
simple exponential spatial decay in the direction of propagation (shown on the log
scale in Fig. 37b). We define the decay length ξ as the distance over which the wave
amplitude decreases by a factor of e. By fitting the data of Fig. 37b with the function
exp(−2y/ξ), we obtain ξ = 1.3 μm. We note that this value is close to the best spin-
wave propagation characteristics obtained in low-loss Py films with comparable
thickness, despite the higher dynamical losses expected due to the stronger damping
in CoFe.
By analyzing transverse cross-sections of the BLS intensity map (Fig. 37c),
we determine the transverse full width at half maximum w of the spin wave
intensity distribution for different positions along the waveguide. The obtained value
w = 320 nm is independent of the propagation coordinate (Fig. 37d), which confirms
that the spin wave is efficiently localized in the waveguide without spreading out.
We note that the measured spatial profile (Fig. 37c) represents a convolution of
the actual profile of the spin wave intensity with the distribution of intensity in
the diffraction-limited probing light spot whose estimated diameter is 250 nm.
The value w = 320 nm is therefore in a reasonable agreement with the measured
waveguide width of 200 nm (inset in Fig. 36a).

Conclusion and Outlook

The post-CMOS information technology will require radically new solutions for
digital and analog information processing. One promising approach is to employ the
spin degree of freedom of electron for information storage and computing, which
is the main focus of the rapidly growing field of spintronics [101–104]. In this
new signal-processing paradigm signals will be codes in terms of a spin angular
momentum that can be carried by either polarized electrons or spin waves.
The use of spin waves (magnons) as carriers of spin angular momentum is
preferable to the use of spin-polarized electrons, because Gilbert magnetic damping,
associated with the transport of spin waves, is, usually, lower than the Ohmic losses
associated with the transport of electrons. The typical medium for the spin wave
propagation in the existing nano-scale magnonics is a soft magnetic metal such
as permalloy (Py). This material choice is mainly dictated by the relative ease of
340 S. O. Demokritov and A. N. Slavin

magnetic information readout via various types of magneto-resistance observed in


metallic ferromagnetic heterostructures. A substantial progress has been made in
this field during the last two decades. In particular, generation of self-sustained
microwave magnetic oscillations by STT effect from spin-polarized currents
[27, 29] as well as pure spin currents arising from spin-Hall effect [31, 94, 95, 105]
have been demonstrated. Novel techniques for precise characterization of magneti-
zation dynamics in nano-scale metallic magnetic systems, such as the technique of
spin-torque ferromagnetic resonance (ST-FMR) [106, 107], have been developed.
In spite of the rapid research progress in the field of metal-based magnonics,
several significant limitations of the metal-based magnonic systems are very evident.
One of them is the relatively large magnetic damping of ferromagnetic metals,
which translates into large spin current densities needed to induce magnetization
switching or self-generation of spin waves in ferromagnetic metals. The large
magnetic damping also results in short propagation lengths (typically ∼1 μm)
of magnons in metallic magnetic nanostructures, which critically hinders the
transition from single magnonic elements to large-scale spintronic circuits based on
propagating magnons. In addition, high electric conductivity of metallic magnets
and the corresponding short charge screening length do not allow to employ
magneto-electric effects such as the recently predicted flexoelectric effect [108–110]
for manipulation of spin waves using electric field.
These drawbacks are absent in magnetic dielectrics, the most common of
which is yttrium iron garnet (YIG) – a ferrimagnetic insulator with very low
magnetic damping (magnon lifetimes reaching 1 μs and magnon propagation length
exceeding 1 cm) [35]. However, the technique of liquid phase epitaxy typically
employed for the growth of high-quality YIG crystals does not allow for deposition
of films sufficiently thin for observation of interfacial spin-dependent phenomena,
which will determine the future of manipulation of spin waves at nanoscale.
The pioneering experiments in YIG-based spintronics performed on the relatively
thick (∼1–3 μm) epitaxial YIG films revealed some weak effects, but failed to
demonstrate reproducible excitation or/and manipulation of propagating magnons
by interfacial spintronic effects [12, 13, 111, 112].
The recently developed methods for growth of ultra-thin (∼ 10 nm) high-quality
YIG films (ferromagnetic resonance (FMR) linewidth ∼3–5 Oe) by pulsed laser
deposition (PLD) [113–115] and patterning of thin YIG films [116] remove major
roadblocks for using magnetic dielectrics in nano-scale spintronic devices and
open a new field of magnon-based spintronics of magnetic dielectrics. Pioneering
experiments in this field performed in the last two years demonstrated excitation
of magnonic signals in magnetic dielectrics by interfacial spin orbit torques,
compensation of magnetic damping in magnetic dielectrics by pure spin currents,
and resulting self-sustained generation of microwave magnetization oscillations in
YIG film samples [117, 118].
We firmly believe that spin waves propagating in magnetic dielectrics and
antiferromagnets will determine the future of the microwave signal processing at
nanoscale, and will form a basis for a new generation of energy-efficient microwave
signal processing devices which will use spin-orbital effects, like the spin-Hall
6 Spin Waves 341

effect [9–11] and inverse spin-Hall effect [119], and electric fields (e.g., through
the flexoelectric effect [109, 110]) for generation, reception, and manipulation of
signals coded in terms of the spin angular momentum and carried by different spin-
wave modes in magnetic nanostructures.

References
1. Bloch, F.: Zur Theorie des Ferromagnetismus. Z. Phys. 61, 206–219 (1930)
2. Neusser, S., Grundler, D.: Magnonics: spin waves on the nanoscale. Adv. Mater. 21, 2927–
2932 (2009)
3. Kruglyak, V.V., Demokritov, S.O., Grundler, D.: Magnonics. J. Phys. D. Appl. Phys. 43,
264001 (2010)
4. Lenk, B., Ulrichs, H., Garbs, F., Münzenberg, M.: The building blocks of magnonics. Phys.
Rep. 507, 107 (2011)
5. Chumak, A.V., Vasyuchka, V.I., Serga, A.A., Hillebrands, B.: Nat. Phys. 11, 453–461 (2015)
6. Slonczewski, J.C.: Current-driven excitation of magnetic multilayers. J. Magn. Magn. Mater.
159, L1–L7 (1996)
7. Berger, L.: Emission of spin waves by a magnetic multilayer traversed by a current. Phys.
Rev. B. 54, 9353–9358 (1996)
8. Slonczewski, J.C.: Excitation of spin waves by an electric current. J. Magn. Magn. Mater.
195, L261–L268 (1999)
9. Dyakonov, M.I., Perel, V.I.: Possibility of orienting electron spins with current. Sov. Phys.
JETP Lett. 13, 467–469 (1971)
10. Hirsch, J.E.: Spin Hall effect. Phys. Rev. Lett. 83, 1834–1837 (1999)
11. Hoffmann, A.: Spin Hall effects in metals. IEEE Trans. Magn. 49, 5172–5193 (2013)
12. Kajiwara, Y., et al.: Transmission of electrical signals by spin-wave interconversion in a
magnetic insulator. Nature. 464, 262–266 (2010)
13. Wang, Z., Sun, Y., Wu, M., Tiberkevich, V., Slavin, A.: Control of spin waves in a
ferromagnetic insulator through interfacial spin scattering. Phys. Rev. Lett. 107, 146602
(2011)
14. Demidov, V.E., Urazhdin, S., Rinkevich, A.B., Reiss, G., Demokritov, S.O.: Spin Hall
controlled magnonic microwaveguides. Appl. Phys. Lett. 104, 152402 (2014)
15. An, K., et al.: Control of propagating spin waves via spin transfer torque in a metallic bilayer
waveguide. Phys. Rev. B. 89, 140405(R) (2014)
16. Urazhdin, S., Demidov, V.E., Ulrichs, H., Kendziorczyk, T., Kuhn, T., Leuthold, J., Wilde, G.,
Demokritov, S.O.: Nanomagnonic devices based on the spin-transfer torque. Nat Nanotech-
nol. 9, 509–513 (2014)
17. Kittel, C.: Excitation of spin waves in a ferromagnet by uniform RF field. Phys. Rev. 110,
1295 (1958)
18. Seavey Jr., M.H., Tannenwald, E.: Direct observation of spin wave resonance. Phys. Rev. Lett.
1, 168 (1958)
19. Damon, R.W., Eshbach, J.R.: Magnetostatic modes of a ferromagnet slab. J. Phys. Chem.
Solids. 19, 308–320 (1961)
20. Grünberg, P., Metawe, F.: Light scattering from bulk and surface spin waves in EuO. Phys.
Rev. Lett. 39, 1561 (1977)
21. Grünberg, P., Schreiber, R., Pang, Y., Brodsky, M.B., Sowers, H.: Layered magnetic struc-
tures: evidence for antiferromagnetic coupling of Fe layers across Cr interlayers. Phys. Rev.
Lett. 57, 2442 (1986)
22. Mathieu, C., Jorzick, J., Frank, A., Demokritov, S.O., Slavin, A.N., Hillebrands, B., Barten-
lian, B., Chappert, C., Decanini, D., Rousseaux, F., Cambrill, E.: Lateral quantization of spin
waves in micron size magnetic wires. Phys. Rev. Lett. 81, 3968 (1998)
342 S. O. Demokritov and A. N. Slavin

23. Jorzick, J., Demokritov, S.O., Hillebrands, B., Berkov, D., Gorn, N.L., Guslienko, K., Slavin,
A.N.: Spin wave wells in nonellipsoidal micrometer size magnetic elements. Phys. Rev. Lett.
88, 047204 (2002)
24. Park, J., Eames, D.M., Engebretson, J., Berezovsky, A., Crowell, P.: Phys. Rev. Lett. 89,
277201 (2002)
25. Tsoi, M., et al.: Excitation of a magnetic multilayer by an electric current. Phys. Rev. Lett.
80, 4281–4284 (1998)
26. Tsoi, M., Jansen, A.G.M., Bass, J., Chiang, W.C., Tsoi, V., Wyder: Generation and detection
of phase-coherent current-driven magnons in magnetic multilayers. Nature. 406, 46 (2000)
27. Kiselev, S.I., et al.: Microwave oscillations of a nanomagnet driven by a spin-polarized
current. Nature. 425, 380–383 (2003)
28. Rippard, W.H., Pufall, M.R., Kaka, S., Russek, S.E., Silva, T.J.: Direct-current induced
dynamics in Co90Fe10/Ni80Fe20 point contacts. Phys. Rev. Lett. 92, 027201 (2004)
29. Krivorotov, I.N., Emley, N.C., Sankey, J.C., Kiselev, S.I., Ralph, D.C., Buhrman, R.A.: Time-
domain measurements of nanomagnet dynamics driven by spin-transfer torques. Science. 307,
228 (2005)
30. Demidov, V.E., Urazhdin, S., Demokritov, S.O.: Direct observation and mapping of spin
waves emitted by spin-torque nano-oscillators. Nat. Mater. 9, 984–988 (2010)
31. Demidov, V.E., et al.: Magnetic nano-oscillator driven by pure spin current. Nat. Mater. 11,
1028–1031 (2012)
32. Landau, L.D., Lifshitz, E.M.: Theory of the dispersion of magnetic permeability in ferromag-
netic bodies. Phys. Z. Sowjet. 8, 153 (1935)
33. L’vov, V.S.: Wave Turbulence Under Parametric Excitation. Springer, New York (1994)
34. Cottam, M.G. (ed.): Linear and Nonlinear Spin Waves in Magnetic Films and Superlattices.
World Scientific, Singapore (1994)
35. Gurevich, A.G., Melkov, G.A.: Magnetization Oscillations and Waves. CRC, New York
(1996)
36. Herring, C., Kittel, C.: On the theory of spin waves in ferromagnetic media. Phys. Rev. 81,
869 (1951)
37. Kalinikos, B.A., Slavin, A.N.: Theory of dipole-exchange spin wave spectrum for ferromag-
netic films with mixed exchange boundary conditions. J. Phys. C. 19, 7013 (1986)
38. Rado, G.T., Weertman, J.R.: Spin-wave resonance in a ferromagnetic metal. J. Phys. Chem.
Solids. 11, 315–333 (1959)
39. Guslienko, K.Y., Demokritov, S.O., Hillebrands, B., Slavin, A.N.: Dynamic pinning of dipolar
origin in nonellipspoidal magnetic stripes. Phys. Rev. B. 66, 132402 (2002)
40. Guslienko, K.Y., Slavin, A.N.: Boundary conditions for magnetization in magnetic nanoele-
ments. Phys. Rev. B. 72, 014463 (2005)
41. Heinrich, B., Urban, R., Woltersdorf, G.: J. Appl. Phys. 91, 7523 (2002)
42. Hiebert, W.K., Stankiewicz, A., Freeman, M.R.: Phys. Rev. Lett. 79, 1134 (1997)
43. Barman, A., Kruglyak, V.V., Hicken, R.J., Rowe, J.M., Kundrotaite, A., Scott, J., Rahman,
M.: Imaging the dephasing of spin wave modes in a square thin film magnetic element. Phys.
Rev. B. 69, 174426 (2004)
44. Acremann, Y., Back, C.H., Buess, M., Portmann, O., Vaterlaus, A., Pescia, D., Melchior, H.:
Imaging precessional motion of the magnetization vector. Science. 290, 492 (2000)
45. Demokritov, S.O., Hillebrands, B., Slavin, A.N.: Brillouin light scattering studies of confined
spin waves: linear and nonlinear confinement. Phys. Rep. 348, 441 (2001)
46. Demokritov, S.O., Demidov, V.E.: Micro-Brillouin light scattering spectroscopy of magnetic
nanostructures. IEEE Trans. Magn. Adv. Magn. 44, 6 (2008)
47. Demidov, V.E., Demokritov, S.O.: Magnonic waveguides studied by microfocus Brillouin
light scattering. IEEE Trans. Magn. 51, 8578 (2015)
48. Jorzick, J., Demokritov, S.O., Mathieu, C., Hillebrands, B., Bartenlian, B., Chappert, C.,
Rousseaux, F., Slavin, A.: Brillouin light scattering from quantized spin waves in micron-
size magnetic wires. Phys. Rev. B. 60, 15194 (1999)
49. Joseph, R.I., Schlomann, E.: Demagnetizing field in nonellipsoidal bodies. J. Appl. Phys. 36,
1579–1593 (1965)
6 Spin Waves 343

50. Bryant, H.: Suhl: thin-film magnetic patterns in an external field. Appl. Phys. Lett. 54, 2224
(1989)
51. Bayer, C., Demokritov, S.O., Hillebrands, B., Slavin, A.N.: Spin wave wells with multiple
states created in small magnetic elements. Appl. Phys. Lett. 82, 607 (2003)
52. Jersch, J., Demidov, V.E., Fuchs, H., Rott, K., Krzysteczko, P., Munchenberger, J., Reiss, G.,
Demokritov, S.O.: Mapping of localized spin-wave excitations by near-field Brillouin light
scattering. Appl. Phys. Lett. 97, 152502 (2010)
53. Serga, A.A., Schneider, T., Hillebrands, B., Demokritov, S.O., Kostylev, M.: Phase-sensitive
Brillouin light scattering spectroscopy from spin-wave packets. Appl. Phys. Lett. 89, 063506
(2006)
54. Vogt, K., Schultheiss, H., Hermsdoerfer, S.J., Pirro, Serga, A.A., Hillebrands, B.: All-optical
detection of phase fronts of propagating spin waves in a Ni81Fe19 microstripe. Appl. Phys.
Lett. 95, 182508 (2009)
55. Demidov, V.E., Urazhdin, S., Demokritov, S.O.: Control of spin-wave phase and wavelength
by electric current on the microscopic scale. Appl. Phys. Lett. 95, 262509 (2009)
56. Demidov, V.E., Demokritov, S.O., Rott, K., Krzysteczko, J., Reiss, G.: Self-focusing of spin
waves in Permalloy microstripes. Appl. Phys. Lett. 91, 252504 (2007)
57. Khitun, A., Bao, M., Wang, K.L.: Magnonic logic circuits. J. Phys. D. Appl. Phys. 43, 264005
(2010)
58. Demidov, V.E., Jersch, J., Demokritov, S.O., Rott, K., Krzysteczko, J., Reiss, G.: Transforma-
tion of propagating spin-wave modes in microscopic waveguides with variable width. Phys.
Rev. B. 79, 054417 (2009)
59. Demidov, V.E., Kostylev, M., Rott, K., Münchenberger, J., Reiss, G., Demokritov, S.O.:
Excitation of short-wavelength spin waves in magnonic waveguides. Appl. Phys. Lett. 99,
082507 (2011)
60. Demidov, V.E., Demokritov, S.O., Birt, D., O’Gorman, B., Tsoi, M., Li, X.: Radiation of
spin waves from the open end of a microscopic magnetic-film waveguide. Phys. Rev. B. 80,
014429 (2009)
61. Schneider, T., et al.: Nondiffractive subwavelength wave beams in a medium with externally
controlled anisotropy. Phys. Rev. Lett. 104, 197203 (2010)
62. Ulrichs, H., Demidov, V.E., Demokritov, S.O., Urazhdin, S.: Spin-torque nano-emitters for
magnonic applications. Appl. Phys. Lett. 100, 162406 (2012)
63. Adam, J.D.: Analog signal-processing with microwave magnetic. Proc. IEEE. 76, 159–170
(1988)
64. Ishak, W.S.: Magnetostatic wave technology: a review. Proc. IEEE. 76, 171–187 (1988)
65. Demidov, V.E., Kostylev, M., Rott, K., Krzysteczko, J., Reiss, G., Demokritov, S.O.:
Excitation of microwaveguide modes by a stripe antenna. Appl. Phys. Lett. 95, 112509
(2009)
66. Myers, E., Ralph, D.C., Katine, J.A., Louie, R.N., Buhrman, R.A.: Current-induced switching
of domains in magnetic multilayer devices. Science. 285, 867–870 (1999)
67. Katine, J.A., Albert, F.J., Buhrman, R.A., Myers, E.B., Ralph, D.C.: Current-driven magneti-
zation reversal and spin-wave excitations in Co/Cu/Co pillars. Phys. Rev. Lett. 84, 3149–3152
(2000)
68. Ralph, D.C., Stiles, M.D.: Spin transfer torques. J. Magn. Magn. Mater. 320, 1190–1216
(2008)
69. Silva, T.J., Rippard, W.H.: Developments in nano-oscillators based upon spin-transfer point-
contact devices. J. Magn. Magn. Mater. 320, 1260–1271 (2008)
70. Brataas, A., Kent, A.D., Ohno, H.: Current-induced torques in magnetic materials. Nat. Mater.
11, 372 (2012)
71. Locatelli, N., Cros, V., Grollier, J.: Spin-torque building blocks. Nat. Mater. 13, 11 (2014)
72. Chen, T., Dumas, R.K., Eklund, A., Muduli, K., Houshang, A., Awad, A.A., Duerrenfeld,
Malm, B.G., Rusu, A., Akerman, J.: Spin-torque and spin-Hall nano-oscillators. IEEE Trans.
Magn. 99, 1 (2016)
73. Madami, M., et al.: Direct observation of a propagating spin wave induced by spin-transfer
torque. Nat. Nanotechnol. 6, 635–638 (2011)
344 S. O. Demokritov and A. N. Slavin

74. Houshang, A., Iacocca, E., Duerrenfeld, P., Sani, S.R., Akerman, J., Dumas, R.K.: Spin-wave-
beam driven synchronization of nanocontact spin-torque oscillators. Nat. Nanotechnol. 11,
280–286 (2016)
75. Demidov, V.E., Urazhdin, S., Liu, R., Divinskiy, B., Telegin, A., Demokritov, S.O.: Excitation
of coherent propagating spin waves by pure spin currents. Nat. Commun. 7, 10446 (2016)
76. Xiao, J., Bauer, G.E.W.: Spin-wave excitation in magnetic insulators by spin-transfer torque.
Phys. Rev. Lett. 108, 217204 (2012)
77. Jungwirth, T., Wunderlich, J., Olejnik, K.: Spin Hall effect devices. Nat. Mater. 11, 382 (2012)
78. Ando, K., Takahashi, S., Harii, K., Sasage, K., Ieda, J., Maekawa, S., Saitoh, E.: Electric
manipulation of spin relaxation using the spin Hall effect. Phys. Rev. Lett. 101, 036601 (2008)
79. Evelt, M., et al.: High-efficiency control of spin-wave propagation in ultra-thin Yttrium Iron
Garnet by the spin-orbit torque. Appl. Phys. Lett. 108, 172406 (2016)
80. Demidov, V.E., Urazhdin, S., Tiberkevich, V., Slavin, A., Demokritov, S.O.: Control of spin-
wave emission from spin-torque nano-oscillators by microwave pumping. Phys. Rev. B. 83,
060406 (R) (2011)
81. Demidov, V.E., et al.: Control of magnetic fluctuations by spin current. Phys. Rev. Lett. 107,
107204 (2011)
82. Slavin, A., Tiberkevich, V.: Spin wave mode excited by spin-polarized current in a magnetic
nanocontact is a standing self-localized wave bullet. Phys. Rev. Lett. 95, 237201 (2005)
83. Rippard, W.H., Pufall, M.R., Silva, T.J.: Quantitative studies of spin-momentum-transfer-
induced excitations in Co/Cu multilayer films using point-contact spectroscopy. Appl. Phys.
Lett. 82, 1260 (2003)
84. Silberberg, Y.: Collapse of optical pulses. Opt. Lett. 15, 1282 (1990)
85. Rezende, S.M., de Aguiar, F.M., Azevedo, A.: Spin-wave theory for the dynamics induced by
direct currents in magnetic multilayers. Phys. Rev. Lett. 94, 037202 (2005)
86. Slavin, A.N., Kabos: Approximate theory of microwave generation in a current-driven
magnetic nanocontact magnetized in an arbitrary direction. IEEE Trans. Magn. 41, 1264
(2005)
87. Akhmediev, N.N., Ankiewicz, A.: Solitons. Nonlinear Pulses and Beams. Chapman & Hall,
London (1997)
88. Aranson, I.S., Kramer, L.: The world of the complex Ginzburg-Landau equation. Rev. Mod.
Phys. 74, 99 (2002)
89. Slavin, A.N., Tiberkevich, V.S.: Excitation of spin waves by spin-polarized current in
magnetic nano-structures. IEEE Trans. Magn. 44, 1916–1927 (2008)
90. Slavin, A.N., Tiberkevich, V.S.: Nonlinear auto-oscillator theory of microwave generation by
spin-polarized current. IEEE Trans. Magn. 45, 1875 (2009)
91. Gerhart, G., Bankowski, E., Melkov, G.A., Tiberkevich, V.S., Slavin, A.N.: Angular depen-
dence of the microwave-generation threshold in a nanoscale spin-torque oscillator. Phys. Rev.
B. 76, 024437 (2007)
92. Consolo, G., Azzerboni, B., Lopez-Diaz, L., Gerhart, G., Bankowski, E., Tiberkevich, V.,
Slavin, A.N.: Micromagnetic study of the above-threshold generation regime in a spin-torque
oscillator based on a magnetic nanocontact magnetized at an arbitrary angle. Phys. Rev. B.
78, 014420 (2008)
93. Bonetti, S., et al.: Experimental evidence of self-localized and propagating spin wave modes
in obliquely magnetized current-driven nanocontacts. Phys. Rev. Lett. 105, 217204 (2010)
94. Liu, R.H., Lim, W.L., Urazhdin, S.: Spectral characteristics of the microwave emission by the
spin Hall nano-oscillator. Phys. Rev. Lett. 110, 147601 (2013)
95. Duan, Z., et al.: Nanowire spin torque oscillator driven by spin orbit torques. Nat. Commun.
5, 5616 (2014)
96. Hoefer, M.A., Silva, T.J., Keller, M.W.: Theory for a dissipative droplet soliton excited by a
spin torque nanocontact. Phys. Rev. B. 82, 054432 (2010)
97. Mohseni, S.M., Sani, S.R., Persson, J., Nguyen, T.N.A., Chung, S., Pogoryelov, Y., Muduli,
K., Iacocca, E., Eklund, A., Dumas, R.K., Bonetti, S., Deac, A., Hoefer, M.A., Akerman, J.:
Spin torque–generated magnetic droplet solitons. Science. 339, 1295 (2013)
6 Spin Waves 345

98. Chung, S., Mohseni, M., Sani, S.R., Iacocca, E., Dumas, R.K., Anh Nguyen, T.N., Pogo-
ryelov, Y., Muduli, K., Eklund, A., Hoefer, M., Akerman, J.: Spin transfer torque generated
magnetic droplet solitons. J. Appl. Phys. 115, 172612 (2014)
99. Ivanov, B.A., Kosevich, A.M.: Zh. Eksp. Teor. Fiz. 72, 2000 (1977)
100. Kosevich, A.M., Ivanov, B.A., Kovalev, A.S.: Magnetic solitons. Phys. Rep. 194, 117–238
(1990)
101. Wolf, S.A., Awschalom, D.D., Buhrman, R.A., Daughton, J.M., von Molnar, S., Roukes,
M.L., Chtchelkanova, A.Y., Treger, D.M.: Spintronics: a spin-based electronics vision for
the future. Science. 294, 1488 (2001)
102. Zutic, I., Fabian, J., Das Sarma, D.: Spintronics: fundamentals and applications. Rev. Mod.
Phys. 76, 323 (2004)
103. Kang, W., Zhang, Y., Wang, Z.H., Klein, J.O., Chappert, C., Ravelosona, D., Wang, G.F.,
Zhang, Y.G., Zhao, W.S.: Spintronics: emerging ultra-low-power circuits and systems beyond
MOS technology. ACM J. Emerg. Technol. Comput. Syst. 12, 16 (2015)
104. Hoffmann, A., Bader, S.D.: Opportunities at the frontiers of spintronics. Phys. Rev. Appl. 4,
047001 (2015)
105. Liu, L.Q., Pai, C.F., Ralph, D.C., Buhrman, R.A.: Magnetic oscillations driven by the spin
Hall effect in 3-terminal magnetic tunnel junction devices. Phys. Rev. Lett. 109, 186602
(2012)
106. Tulapurkar, A.A., Suzuki, Y., Fukushima, A., Kubota, H., Maehara, H., Tsunekawa, K.,
Djayaprawira, D.D., Watanabe, N., Yuasa, S.: Spin-torque diode effect in magnetic tunnel
junctions. Nature. 438, 339 (2005)
107. Sankey, J.C., Braganca, M., Garcia, A.G.F., Krivorotov, I.N., Buhrman, R.A., Ralph, D.C.:
Spin-transfer-driven ferromagnetic resonance of individual nanomagnets. Phys. Rev. Lett. 96,
227601 (2006)
108. Bar’yakhtar, V.G., L’vov, V.A., Yablonskii, D.A.: Inhomogeneous magnetoelectric effect.
JETP Lett. 37, 673 (1983)
109. Dzyaloshinskii, I.: Magnetoelectricity in ferromagnets. Europhys. Lett. 83, 67001 (2008)
110. Mills, D.L., Dzyaloshinskii, I.E.: Influence of electric fields on spin waves in simple
ferromagnets: role of the flexoelectric interaction. Phys. Rev. B. 78, 184422 (2008)
111. Sandweg, C.W., Kajiwara, Y., Chumak, A.V., Serga, A.A., Vasyuchka, V.I., Jungfleisch, M.B.,
Saitoh, E., Hillebrands, B.: Spin pumping by parametrically excited exchange magnons. Phys.
Rev. Lett. 106, 216601 (2011)
112. Chumak, A.V., Serga, A.A., Jungfleisch, M.B., Neb, R., Bozhko, D.A., Tiberkevich, V.S.,
Hillebrands, B.: Direct detection of magnon spin transport by the inverse spin Hall effect.
Appl. Phys. Lett. 100, 082405 (2012)
113. Heinrich, B., Burrowes, C., Montoya, E., Kardasz, B., Girt, E., Song, Y.-Y., Sun, Y., Wu, M.:
Spin pumping at the magnetic insulator (YIG)/normal metal (Au) interfaces. Phys. Rev. Lett.
107, 066604 (2011)
114. Sun, Y., Song, Y.-Y., Chang, H., Kabatek, M., Jantz, M., Schneider, W., Wu, M., Schultheiss,
H., Hoffmann, A.: Growth and ferromagnetic resonance properties of nanometer-thick yttrium
iron garnet films. Appl. Phys. Lett. 101, 152405 (2012)
115. Chang, H., Li, Zhang, W., Liu, T., Hoffmann, A., Deng, L., Wu, M.: Nanometer-thick yttrium
iron garnet films with extremely low damping. IEEE Magn. Lett. 5, 6700104 (2014)
116. Hahn, C., Naletov, V.V., de Loubens, G., Klein, O., d’Allivy Kelly, O., Anane, A., Bernard,
R., Jacquet, E., Bortolotti, Cros, V., Prieto, J.L., Munoz, M.: Measurement of the intrinsic
damping constant in individual nanodisks of Y3 Fe5 O12 and Y3 Fe5 O12 |Pt. Appl. Phys. Lett.
104, 152410 (2014)
117. Hamadeh, A., et al.: Electronic control of the spin-wave damping in a magnetic insulator.
Phys. Rev. Lett. 113, 197203 (2014)
118. Collet, M., et al.: Generation of coherent spin-wave modes in Yttrium Iron Garnet microdiscs
by spin-orbit torque. Nat. Commun. 7, 10377 (2016)
119. Saitoh, E., Ueda, M., Miyajima, H., Tatara, G.: Conversion of spin current into charge current
at room temperature: inverse spin-Hall effect. Appl. Phys. Lett. 88, 182509 (2006)
346 S. O. Demokritov and A. N. Slavin

Sergej O. Demokritov received his Ph.D at Kapitsa Institute for


Physical Problems, Moscow, Russia. In the 1990s, he moved to
Germany to start to work with P. Grünberg at Research Center
Jülich. Since 2004, he is a Professor at Münster University, Ger-
many. His main directions of research are dynamics and quantum
thermodynamics of magnetic structures, spin-wave research, and
magnonics.

Andrei N. Slavin is a Distinguished Professor and Chair of the


Physics Department, Oakland University, Michigan, USA. He
received his Ph.D from the St. Petersburg Technical University,
Russia. Andrei is Fellow of the American Physical Society and
Fellow of the IEEE. He is a specialist in magnetization dynamics
and spin waves and published over 280 research papers in this
field.
Micromagnetism
7
Lukas Exl , Dieter Suess , and Thomas Schrefl

Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 348
Micromagnetics Basics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 349
Magnetic Gibbs Free Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 350
Spin, Magnetic Moment, and Magnetization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 350
Exchange Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 351
Magnetostatics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 356
Zeeman Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 357
Magnetostatic Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 357
Crystal Anisotropy Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 363
Magnetoelastic and Magnetostrictive Energy Terms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 366
Characteristic Length Scales . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 372
Exchange Length . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 372
Critical Diameter for Uniform Rotation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 373
Wall Parameter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 375
Mesh Size in Micromagnetic Simulations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 376

L. Exl ()
University of Vienna Research Platform MMM Mathematics – Magnetism – Materials,
University of Vienna, and Wolfgang Pauli Institute, Wien, Austria
e-mail: lukas.exl@univie.ac.at
D. Suess
University of Vienna Research Platform MMM Mathematics – Magnetism – Materials, and
Physics of Functional Materials, Faculty of Physics, University of Vienna, Wien, Austria
e-mail: dieter.suess@univie.ac.at
T. Schrefl
Christian Doppler Laboratory for Magnet Design Through Physics Informed Machine Learning,
Department of Integrated Sensor Systems, Danube University Krems, Wiener Neustadt, Austria
e-mail: thomas.schrefl@donau-uni.ac.at

© Springer Nature Switzerland AG 2021 347


J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_7
348 L. Exl et al.

Brown’s Micromagnetic Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 376


Euler Method: Finite Differences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 377
Ritz Method: Finite Elements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 380
Magnetization Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 382
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 386

Abstract

Computational micromagnetics is widely used for the design and development


of magnetic devices. The theoretical background of these simulations is the
continuum theory of micromagnetism. It treats magnetization processes on a
significant length scale which is small enough to resolve magnetic domain walls
and large enough to replace atomic spins by a continuous function of position.
The continuous expression for the micromagnetic energy terms are either
derived from their atomistic counterpart or result from symmetry arguments.
The equilibrium conditions for the magnetization and the equation of motion
are introduced. The focus of the discussion lies on the equations that form the
basic building blocks of micromagnetic solvers. Numerical examples illustrate
the micromagnetic concepts. An open-source simulation environment was used
to address the ground state of thin film magnetic elements, initial magnetization
curves, stress-driven switching of magnetic storage elements, the grain size
dependence of the coercivity of permanent magnets, and damped oscillations
in magnetization dynamics.

Introduction

Computer simulations are essential tools for product design in modern society. This
is also true for magnetic materials and their applications. The design of magnetic
data storage systems such as hard disk devices [1, 2, 3, 4, 5] and random access
memories [6, 7] relies heavily on computer simulations. Similarly, the computer
models assist the development of magnetic sensors [8, 9] used as biosensors or
position and speed sensors in automotive applications [10]. Computer simulations
give guidance for the advance of high performance permanent magnet materials
[11, 12, 13] and devices. In storage and sensor applications, the selection of
magnetic materials, the geometry of the magnetically active layers, and the layout
of current lines are key design questions that can be answered by computations. In
addition to the intrinsic magnetic properties, the microstructure including grain size,
grain shape, and grain boundary phases is decisive for the magnet’s performance.
Computer simulations can quantify the influence of microstructural features on the
remanence and the coercive field of permanent magnets.
The characteristic length scale of the abovementioned computer models is
in the range of nanometers to micrometers. The length scale is too big for a
description by spin polarized density functional theory. Efficient simulations by
atomistic spin dynamics [14] are possible for nano-scale devices only. On the other
hand, macroscopic simulations using Maxwell’s equations hide the magnetization
7 Micromagnetism 349

processes that are relevant for the specific functions of the material or device under
consideration. Micromagnetism is a continuum theory that describes magnetization
processes on significant length scales that are

• large enough to replace discrete atomic spins by a continuous function of position


(the magnetization), but
• small enough to resolve the transition of the magnetization between magnetic
domains

For most ferromagnetic materials, this length scale is in the range of a few
nanometers to micrometers. The first aspect leads to a mathematical formulation
which makes it possible to simulate materials and devices in a reasonable time.
Instead of billions of atomic spins, only millions of finite elements have to be taken
into account. The second aspect keeps all relevant physics so that the influence of
structure and geometry on the formation of reversed domains and the motion of
domain walls can be computed.
The theory of micromagnetism was developed well before the advance of modern
computing technology. Key properties of magnetic materials can be understood by
analytic or semi-analytic solutions of the underlying equations. However, the future
use of powerful computers for the calculation of magnetic properties by solving the
micromagnetic equations numerically was already proposed by Brown [15] in the
late 1950s. The purpose of micromagnetics is the calculation of the magnetization
distribution as function of the applied field or the applied current taking into account
the structure of the material and the mutual interactions between the different
magnetic parts of a device.

Micromagnetics Basics

The key assumption of micromagnetism is that the spin direction changes only by
a small angle from one lattice point to the next [16]. The direction angles of the
spins can be approximated by a continuous function of position. Then the state
of a ferromagnet can be described by a continuous vector field, the magnetization
M(x). The magnetization is the magnetic moment per unit volume. The direction of
M(x) varies continuously with the coordinates x, y, and z. Here we introduced the
position vector x = (x, y, z). Starting from the Heisenberg model [17, 18] which
describes a ferromagnet by interacting spins associated with each atom, we derive
the micromagnetic equations whereby several assumptions are made:

1. Micromagnetism is a quasi-classical theory. The spin operators of the Heisenberg


model are replaced by classical vectors.
2. The The length of the magnetization vector is a constant that is uniform over
each material of the ferromagnetic body and only depends on temperature.
3. The temperature is constant in time and in space.
4. The Gibbs free energy of the ferromagnetic body is expressed in terms of the
direction cosines of the magnetization.
350 L. Exl et al.

5. The energy terms are derived either by the transition from an atomistic model to
a continuum model or phenomenologically.

In classical micromagnetism, the magnetization can only rotate. A change of the


length of M is forbidden. Thus, a ferromagnet is in thermodynamic equilibrium,
when the torque on the magnetic moment MdV in any volume element dV is zero.
The torque on the magnetic moment MdV caused by a magnetic field H is

T = μ0 MdV × H , (1)

where μ0 is the permeability of vacuum (μ0 = 4π × 10−7 Tm/A). The equilibrium


condition (1) follows from the direct variation of the Gibbs free energy. If only the
Zeeman energy of the magnet in an external field is considered, H is the external
field, H ext . In general additional energy terms will be relevant. Then H has to be
replaced by the effective field, H eff . Each energy term contributes to the effective
field.
In section “Magnetic Gibbs Free Energy”, we will derive continuum expressions
for the various contributions to the Gibbs free energy functional using the direction
cosines of the magnetization as unknown functions. In section “Characteristic
Length Scales”, we discuss the different characteristic length scales used to describe
magnetic phenomena. In section “Brown’s Micromagnetic Equation”, we show
how the equilibrium condition can be obtained by direct variation of the Gibbs free
energy functional.

Magnetic Gibbs Free Energy

We describe the state of the magnet in terms of the magnetization M(x). In the
following we will show how the continuous vector field M(x) is related to the
magnetic moments located at the atom positions of the magnet.

Spin, Magnetic Moment, and Magnetization

The local magnetic moment of an atom or ion at position x i is associated with the
spin angular momentum, h̄S,

|e|
μ(x i ) = −g h̄S(x i ) = −gμB S(x i ). (2)
2m

Here e is the charge of the electron, m is the electron mass, and g is the Landé factor.
The Landé factor is g ≈ 2 for metal systems with quenched orbital moment. The
constant μB = 9.274 × 10−24 Am2 = 9.274 × 10−24 J/T is the Bohr magneton.
The constant h̄ is the reduced Planck constant, h̄ = h/(2π ), where h is the Planck
constant. The magnetization of a magnetic material with N atoms per unit volume is
7 Micromagnetism 351

M = Nμ. (3)

The magnetic moment is often given in Bohr magnetons per atom or Bohr
magnetons per formula unit. The magnetization is

M = Nfu μfu , (4)

where μfu is the magnetic moment per formula unit and Nfu is the number of
formula units per unit volume.
The length of the magnetization vector is assumed to be a function of temperature
only and does not depend on the strength of the magnetic field:

|M| = Ms (T ) = Ms , (5)

where Ms is the saturation magnetization. In classical micromagnetism the temper-


ature, T, is assumed to be constant over the ferromagnetic body and independent of
time t. Therefore Ms is fixed and time evolution of the magnetization vector can be
expressed in terms of the unit vector m = M/|M|

M(x, t) = m(x, t)Ms . (6)

The saturation magnetization of a material is frequently given as μ0 Ms in units of


Tesla.
Example. The saturation magnetization is an input parameter for micromagnetic
simulations. In a multiscale simulation approach of the hysteresis properties of a
magnetic material, it may be derived from the ab initio calculation of magnetic
moment per formula unit. In NdFe11 TiN, the calculated magnetic moment per
formula unit is 26.84 μB per formula unit [19]. The computed lattice constants
were a = 8.537 × 10−10 m, b = 8.618 × 10−10 m, and c = 4.880 × 10−10 m [19]
which give a volume of the unit cell of v = 359.0×10−30 m3 . There are two formula
units per unit cell and Nfu = 2/v = 5.571 × 1027 . With (4) and (5), the saturation
magnetization of NdFe11 TiN is Ms = 1.387 × 106 A/m (μ0 Ms = 1.743 T).

Exchange Energy

The exchange energy is of quantum mechanical nature. The energy of two ferromag-
netic electrons depends on the relative orientation of their spins. When the two spins
are parallel, the energy is lower than the energy of the antiparallel state. Qualitatively
this behavior can be explained by the Pauli exclusion principle and the electrostatic
Coulomb interaction. Owing to the Pauli exclusion principle, two electrons can only
be at the same place if they have opposite spins. If the spins are parallel, the electrons
tend to move apart which lowers the electrostatic energy. The corresponding gain in
energy can be large enough so that the parallel state is preferred.
352 L. Exl et al.

The exchange energy, Eij , between two localized spins is [18]

Eij = −2Jij S i · S j , (7)

where Jij is the exchange integral between atoms i and j and h̄S i is the angular
momentum of the spin at atom i. For cubic metals and hexagonal closed packed
metals with ideal c over a ratio there holds Jij = J . Treating the exchange energy
for a large number of coupled spins, we regard Eij as a classical potential energy
and replace S i by a classical vector. Let mi be the unit vector in direction −S i . Then
mi is the unit vector of the magnetic moment at atom i. If ϕij is the angle between
the vectors mi and mj , the exchange energy is

Eij = −2J S 2 cos(ϕij ), (8)

where S = |S i | = |S j | is the spin quantum number.


Now, we introduce a continuous unit vector m(x) and assume that the angle
ϕij between the vectors mi and mj is small. We set m(x i ) = mi and expand m
around x i

m(x i + a j ) =m(x i )+
∂m ∂m ∂m
aj + bj + cj +
∂x ∂y ∂z (9)
 
1 ∂ 2m 2 ∂ 2m 2 ∂ 2m 2
a + b + c + ....
2 ∂x 2 j ∂y 2 j ∂z2 j

Here a j = (aj , bj , cj )T is the vector connecting points x i and x j = x i + a j .


We can replace cos(ϕij ) by cos(ϕij ) = m(x i ) · m(x j ) in (8). Summing up
over the six nearest neighbors of a spin in a simple cubic lattice gives (see
Fig. 1) the exchange energy of the unit cell. The vectors a j take the values
(±a, 0, 0)T , (0, ±a, 0)T , (0, 0, ±a)T . For every vector a, there is the correspond-
ing vector −a. Thus the linear terms in the variable a in (9) vanish in the summation.
The same holds for mixed second derivatives in the expansion (9). The constant
term, m · m = 1, plays no role for the variation of the energy and will be neglected.
The exchange energy of a unit cell in a simple cubic lattice is


6 6 
 
∂ 2 mi 2 ∂ 2 mi 2 ∂ 2 mi 2
Eij = − J S 2 mi · a + mi · b + mi · c
∂x 2 j ∂y 2 j ∂z2 j
j =1 j =1
(10)
 
∂ 2 mi ∂ 2 mi ∂ 2 mi
= − 2J S 2 a 2 mi · + mi · + mi ·
∂x 2 ∂y 2 ∂z2

To get the exchange energy of the crystal, we sum over all atoms i and divide by 2
to avoid counting each pair of atoms twice. We also use the relations
7 Micromagnetism 353

Fig. 1 Nearest neighbors of


spin i for the calculation of
the exchange energy in a
simple cubic lattice

 
∂ 2m ∂m 2
m· = − , (11)
∂x 2 ∂x

which follows from differentiating m · m = 1 twice with respect to x. Thus we can


write
      
J S2  3 ∂mi 2 ∂mi 2 ∂mi 2
Eex = a + + . (12)
a ∂x ∂y ∂z
i

The sum in (12) is over the unit cells of the crystal with volume V . In the continuum
limit, we replace the sum with an integral. The exchange energy is

  2  2  2 
∂m ∂m ∂m
Eex = A + + dV . (13)
V ∂x ∂y ∂z

Expanding and rearranging the terms in the bracket and introducing the nabla
operator, ∇, we obtain
 

2
Eex = A (∇mx )2 + ∇my + (∇mz )2 dV . (14)
V
354 L. Exl et al.

In equations (13) and (14), we introduced the exchange constant

J S2
A= n. (15)
a

In cubic lattices, n is the number of atoms per unit cell (n = 1, 2, and 4 for
simple cube, body-centered cubic, and face-centered cubic lattices, respectively).
In a hexagonal
√ closed packed structures, n is the ideal nearest neighbor distance
(n = 2 2). The number N of atoms per unit volume is n/a 3 . At non-zero
temperature, the exchange constant may be expressed in terms of the saturation
magnetization, Ms (T ). Formally we replace S by its thermal average. Using
equations (2) and (3), we rewrite

J [Ms (T )]2
A(T ) = n. (16)
(NgμB )2 a

The calculation of the exchange constant by (15) requires a value for the
exchange integral, J . Experimentally, one can measure a quantity that strongly
depends on J such as the Curie temperature, TC ; the temperature dependence of
the saturation magnetization, Ms (T ); or the spin wave stiffness parameter, in order
to determine J. According to the molecular field theory [20], the exchange integral
is related to the Curie temperature given by

3 k B TC 1 3 k B TC S n
J = or A = . (17)
2 S(S + 1) z 2 a(S + 1) z

The second equation follows from the first one by replacing J with the relation (15).
Here z is the number of nearest neighbors (z = 6, 8, 12, and 12 for simple cubic,
body-centered cubic, face-centered cubic, and hexagonal closed packed lattices,
respectively) and kB =1.3807×10−23 J/K is Boltzmann’s constant. The use of (17)
together with (15) underestimates the exchange constant by more than a factor of 2
[21]. Alternatively one can use the temperature dependence of the magnetization as
arising from the spin wave theory

Ms (T ) = Ms (0)(1 − CT 3/2 ). (18)

Equation (18) is valid for low temperatures. From the measured temperature
dependence Ms (T ), the constant C can be determined. Then the exchange integral
[21, 22] and the exchange constant can be calculated from C as follows:
 2/3  2/3
0.0587 kB 0.0587 kB
J = or A = . (19)
nSC 2S n2 S 2 C 2a

This method was used by Talagala and co-workers [23]. They measured the
temperature dependence of the saturation magnetization in NiCo films to determine
7 Micromagnetism 355

the exchange constant as function of the Co content. The exchange constant can also
be evaluated from the spin wave dispersion relation (see Chapter SPW) which can
be measured by inelastic neutron scattering, ferromagnetic resonance, or Brillouin
light scattering [24]. The exchange integral [22] and the exchange constant are
related to the spin wave stiffness constant, D, via the following relations:

D 1 D
J = 2
or A = NS. (20)
2 Sa 2

For the evaluation of the exchange constant, we can use S = Ms (0)/(NgμB ) [25]
for the spin quantum number in equations (17), (19), and (20). This gives the relation
between the exchange constant, A, and the spin wave stiffness constant, D,

DMs (0)
A= , (21)
2gμB

when applied to (20). Using neutron Brillouin Scattering, Ono and co-workers [26]
measured the spin wave dispersion in a polycrystalline Nd2 Fe14 B magnet, in order
to determine its exchange constant.
Ferromagnetic exchange interactions keep the magnetization uniform. Depend-
ing on the sample, geometry external fields may lead to a locally confined
non-uniform magnetization. Probing the magnetization twist experimentally and
comparing the result with the computed equilibrium magnetic state (see sec-
tion “Brown’s Micromagnetic Equation”) is an alternative method to determine the
exchange constant. The measured data is fitted to the theoretical model whereby
the exchange constant is a free parameter. Smith and co-workers [27] measured
the anisotropic magnetoresistance to probe the fanning of the magnetization in a
thin permalloy film from which its exchange constant was calculated. Eyrich and
co-workers [24] measured the field-dependent magnetization, M(H ), of a trilayer
structure in which two ferromagnetic films are coupled antiferromagnetically. The
M(H ) curve probes the magnetization twist within the two ferromagnets. Using
this method the exchange constant of Co alloyed with various other elements was
measured [24].
The interplay between the effects of ferromagnetic exchange coupling, magneto-
static interactions, and the magnetocrystalline anisotropy leads to the formation of
domain patterns (for details on domain structures, see Chapter Domains). With mag-
netic imaging techniques, the domain width, the orientation of the magnetization,
and the domain wall width can be measured. These values can be also calculated
using a micromagnetic model of the domain structure. By comparing the predicted
values for the domain width with measured data, Livingston [28, 29] estimated
the exchange constant of the hard magnetic materials SmCo5 and Nd2 Fe14 B. This
method can be improved by comparing more than one predicted quantity with
measured data. Newnham and co-workers [30] measured the domain width, the
orientation of the magnetization in the domain, and the domain wall width in foils
356 L. Exl et al.

of Nd2 Fe14 B. By comparing the measured values with the theoretical predictions,
they estimated the exchange constant of Nd2 Fe14 B.
Input for micromagnetic simulations: The high temperature behavior of perma-
nent magnets is of utmost importance for the applications of permanent magnets
in the hybrid or electric vehicles. For computation of the coercive field by
micromagnetic simulations, the exchange constant is needed as input parameter.
Values for A(T ) may be obtained from the room temperature value of A(300 K)
and Ms (T ). Applying (16) gives A(T ) = A(300 K) × [Ms (T )/Ms (300 K)]2 .

Magnetostatics

We now consider the energy of the magnet in an external field produced by


stationary currents and the energy of the magnet in the field produced by the
magnetization of the magnet itself. The latter field is called demagnetizing field.
In micromagnetics, these fields are treated statically if eddy currents are neglected.
In magnetostatics, we have no time-dependent quantities. In the presence of a
stationary magnetic current, Maxwell’s equations reduce to [31]

∇ ×H = j (22)
∇ ·B = 0 (23)

Here B is the magnetic induction or magnetic flux density, H is the magnetic field,
and j is the current density. The charge density fulfills ∇ · j = 0 which expresses
the conservation of electric charge. We now have the freedom to split the magnetic
field into its solenoidal and nonrotational part

H = H ext + H demag . (24)

By definition, we have

∇ · H ext = 0, (25)
∇ × H demag = 0. (26)

Using (22) and (24), we see that the external field, H ext , results from the current
density (Ampere’s law)

∇ × H ext = j . (27)

On a macroscopic length scale, the relation between the magnetic induction and the
magnetic field is expressed by

B = μH , (28)
7 Micromagnetism 357

where μ is the permeability of the material. Equation (28) is used in magnetostatic


field solvers [32] for the design of magnetic circuits. In these simulations, the
permeability describes the response of the material to the magnetic field. Micromag-
netics describes the material on a much finer length scale. In micromagnetics, we
compute the local distribution of the magnetization as function of the magnetic field.
This is the response of the system to (an external) field. Indeed, the permeability
can be derived from micromagnetic simulations [33]. For the calculation of the
demagnetizing field, we can treat the magnetization as fixed function of space.
Instead of (28), we use

B = μ0 (H + M) . (29)

The energy of the magnet in the external field, H ext , is the Zeeman energy. The
energy of the magnet in the demagnetizing field, H demag , is called magnetostatic
energy.

Zeeman Energy

The energy of a magnetic dipole moment, μ, in an external magnetic induction Bext


is −μ · Bext . We use B ext = μ0 H ext and sum over all local magnetic moments at
positions x i of the ferromagnet. The sum,

Eext = −μ0 μi · H ext , (30)
i

is the interaction energy of the magnet with the external field. To obtain the Zeeman
energy in a continuum model, we introduce the magnetization M = Nμ, define the
volume per atom, Vi = 1/N, and replace the sum with an integral. We obtain
 
Eext = −μ0 (M · H ext )Vi → −μ0 (M · H ext )dV . (31)
i V

Using (6), we express the Zeeman energy in terms of the unit vector of the
magnetization 
Eext = − μ0 Ms (m · H ext )dV. (32)
V

Magnetostatic Energy

The magnetostatic energy is also called dipolar interaction energy. In a crystal each
moment creates a dipole field, and each moment is exposed to the magnetic field
created by all other dipoles. Therefore magnetostatic interactions are long range.
The magnetostatic energy cannot be represented as a volume integral over the
magnet of an energy density dependent on only local quantities.
358 L. Exl et al.

Demagnetizing Field as Sum of Dipolar Fields


The total magnetic field at point x i , which is created by all the other magnetic
dipoles, is the sum over the dipole fields from all moments μj = μ(x j )

 
1  (μj · r ij )r ij μj
H dip (x i ) = 3 − 3 . (33)

j =i
rij5 rij

The vectors r ij = x i − x j connect the source points with the field point. The
distance between a source point and a field point is rij = |r ij |. In order to obtain
a continuum expression for the field, we split the sum (33) into two parts. The
contribution to the field from moments that are far from x i will not depend strongly
on their exact position at the atomic level. Therefore we can describe them by a
continuous magnetization and replace the sum with an integral. For moments μj
which are located within a small sphere with radius R around x i , we keep the sum.
Thus we split the dipole field into two parts [34]:

H dip (x i ) = H near (x i ) + H demag (x i ). (34)

Here
 
1  (μj · r ij )r ij μj
H near (x i ) = 3 − (35)

rij <R
rij5 rij3

is the contribution of the sum of the dipoles within the sphere (see Fig. 2). For
the dipoles outside, the sphere we use a continuum approximation. Introducing the
magnetic dipole element MdV  , we can replace the sum in (33) with an integral for
rij ≥ R

Fig. 2 Computation of the


total magnetostatic field at
point atom i. The near field is
evaluated by a direct sum
over all dipoles in the small
sphere. The atomic moments
outside the sphere are
replaced by a continuous
magnetization which
produces the far field acting
on i
7 Micromagnetism 359

 

1 M(x  ) · (x i − x  ) (x i − x  ) M(x  )
H demag (x i ) = 3  |5
−  |3
dV  .
4π V* |x i − x |x i − x
(36)

The integral is over V *, the volume of the magnet without the small sphere around
the field point x i .
The sum in (35) is the contribution of the dipoles inside the sphere to the total
magnetostatic field. The corresponding energy term is local. It can be expressed
as an integral of an energy density that depends only on local quantities [34].
The term depends on the symmetry of the lattice and has the same form as the
crystalline anisotropy. Therefore it is included in the anisotropy energy. When the
anisotropy constants in (58) are determined experimentally, they already include the
contribution owing to dipolar interactions.

Magnetic Scalar Potential


The demagnetizing field is nonrotational. Therefore we can write the demagnetizing
field as gradient of a scalar potential

H demag = −∇U. (37)

Applying −∇ to

1 x − x
U (x) = − M(x  ) · dV  (38)
4π V* |x − x  |3

gives (36). In computational micromagnetics, it is beneficial to work with effective


magnetic volume charges, ρm = −∇  · M(x  ), and effective magnetic surface
charges, σm = M(x  ) · n. Using

x − x 1 1 1
= −∇ and ∇ = −∇  (39)
|x − x  |3 |x − x  | |x − x  | |x − x  |

we obtain

1 1
U (x) = M(x  ) · ∇  dV  . (40)
4π V* |x − x  |

Now we shift the ∇  operator from 1/|x − x  | to M. We use

M(x  ) ∇  · M(x  ) 1
∇ · 
= 
+ M(x  ) · ∇  , (41)
|x − x | |x − x | |x − x  |

apply Gauss’ theorem, and obtain [31]



1 ρm (x  )  1 σm (x  )
U (x) = dV + dS  . (42)
4π V * |x − x  | 4π ∂V * |x − x  |
360 L. Exl et al.

Magnetostatic Energy
For computing the magnetostatic energy, there is no need to take into account
(35). The near field is already included in the crystal anisotropy energy. We now
compute the energy of each magnetic moment μi in the field H demag (x i ) from the
surrounding magnetization. The sum over all atoms is the magnetostatic energy of
the magnet
μ0 
Edemag = − μi · H demag (x i ). (43)
2
i

The factor 1/2 avoids counting each pair of atoms twice. Similar to the procedure
for the exchange and Zeeman energy, we replace the sum with an integral
 
μ0 1

Edemag = − M · H demag = − μ0 Ms m · H demag dV . (44)


2 V 2 V

Alternatively, the magnetostatic energy can be expressed in terms of a magnetic


scalar potential and effective magnetic charges. We start from (44), replace H demag
by −∇U , and apply Gauss’ theorem on ∇ · (MU ) to obtain

μ0 μ0
Edemag = ρm U dV + σm U dS. (45)
2 V 2 ∂V

Equation (45) is widely used in numerical micromagnetics. Its direct variation


(see section “Brown’s Micromagnetic Equation”) with respect to M gives the
cell averaged demagnetizing field. This method was introduced in numerical
micromagnetics by LaBonte [35] and Schabes and Aharoni [36]. For discretization
with piecewise constant magnetization only, the surface integrals remain.
In a uniformly magnetized spheroid, the demagnetizing field is antiparallel to the
magnetization. The demagnetizing field is

H demag = −N M, (46)

where N is the demagnetizing factor. For a sphere the demagnetizing factor is 1/3.
Using (44), we find
μ0
Edemag = NMs2 V (47)
2
for the magnetostatic energy of a uniformly magnetized sphereoid with volume V .
In a cuboid or polyhedral particle, the demagnetizing field is nonuniform. However
we still can apply (47) when we use a volume averaged demagnetizing factor
which is obtained from a volume-averaged demagnetizing field. Interestingly the
volume-averaged demagnetizing factor for a cube is 1/3 the same value as for
the sphere. For a general rectangular prism, Aharoni [37] calculated the volume
averaged demagnetizing factor. A convenient calculation tool for the demagnetizing
7 Micromagnetism 361

factor, which uses Aharoni’s equation, is given on the Magpar website [38]. A
simple approximate equation for the demagnetizing factor of a square prism with
dimensions l × l × pl is [39]

1
N= , (48)
2p + 1

where p is the aspect ratio and N is the demagnetizing factor along the edge with
length pl.

Magnetostatic Boundary Value Problem


Equation (42) is the solution of the magnetostatic boundary value problem, which
can be derived from Maxwell’s equations as follows. From (23) and (29), the
following equation holds for the demagnetizing field

∇ · H demag = −∇ · M. (49)

Plugging (37) into (49), we obtain a partial differential equation for the scalar
potential

∇ 2 U = ∇ · M. (50)

Equation (50) holds inside the magnet. Outside the magnet M = 0 and we have

∇ 2 U = 0. (51)

At the magnet’s boundary, the following interface conditions [31] hold

U (in) = U (out) , (52)



∇U (in) − ∇U (out) · n = M · n, (53)

where n denotes the surface normal. The first condition follows from the continuity
of the component of H parallel to the surface (or ∇ ×H = 0). The second condition
follows from the continuity of the component of B normal to the surface (or ∇ ·B =
0). Assuming that the scalar potential is regular at infinity,

1
|U (x)| ≤ C for |x| large enough and constant C > 0 (54)
|x|

the solution of equations (50) to (53) is given by (42). Formally the integrals in (42)
are over the volume, V *, and the surface, ∂V *, of the magnet without a small sphere
surrounding the field point. The transition from V *→ V adds a term −M/3 to the
field and thus shifts the energy by a constant which is proportional to Ms2 . This is
usually done in micromagnetics [34].
362 L. Exl et al.

The above set of equations for the magnetic scalar potential can also be derived
from a variational principle. Brown [16] introduced an approximate expression
 
 μ0
Edemag = μ0 M · ∇U  dV − (∇U  )2 dV (55)
V 2

for the magnetostatic energy, Edemag . For any magnetization distribution M(x), the
following equation holds


Edemag (M) ≥ Edemag (M, U  ), (56)

where U  is an arbitrary function which is continuous in space and regular at infinity


[16]. A proof of (56) is given by Asselin and Thiele [40]. The inequality (56) is sharp
in the sense that if maximized with respect to the variable U  , equality holds in (56)
and U  is the scalar potential owing to M. Then equality holds and Edemag reduces
to the usual magnetostatic energy Edemag . Equation (55) is used in finite element
micromagnetics for the computation of the magnetic scalar potential. The Euler-
Lagrange equation of (55) with respect to U  gives the magnetostatic boundary value
problem (50) to (53) [40].

Examples
Magnetostatic energy in micromagnetic software: For physicists and software engi-
neers developing micromagnetic software, there are several options to implement
magnetostatic field computation. The choice depends on the discretization scheme,
the numerical methods used, and the hardware. Finite difference solvers including
OOMMF [41], MuMax3 [42], and FIDIMAG [43] use (45) to compute the mag-
netostatic energy and the cell-averaged demagnetizing field. For piecewise constant
magnetization only, the surface integrals over the surfaces of the computational cells
remain. MicroMagnum [44] uses (42) to evaluate the magnetic scalar potential.
The demagnetizing field is computed from the potential by a finite difference
approximation. This method shows a higher speed up on Graphics Processor Units
[45] though its accuracy is slightly less. Finite element solvers compute the magnetic
scalar potential and build its gradient. Magpar [46], Nmag [47], and magnum.fe
[48] solve the partial differential equations (50) to (53). FastMag [49], a finite
element solver, directly integrates (42). Finite difference solvers apply the Fast
Fourier Transforms for the efficient evaluation of the involved convolutions. Finite
element solvers often use hierarchical clustering techniques for the evaluation of
integrals [50].
Magnetic state of nano-elements: From (45), we see that the magnetostatic
energy tends to zero if the effective magnetic charges vanish. This is known as
pole avoidance principle [34]. In large samples where the magnetostatic energy
dominates over the exchange energy, the lowest energy configurations are such that
∇ · M in the volume and M · n on the surface tend to zero. The magnetization
is aligned parallel to the boundary and may show a vortex. These patterns are
known as flux closure states. In small samples, the expense of exchange energy
7 Micromagnetism 363

Fig. 3 Computed magnetization patterns for a soft magnetic square element


(K1 = 0, μ0 Ms = 1 T, A = 10 pJ/m, mesh size h = 0.56 A/(μ0 Ms2 ) = 2 nm) as function of
element size L. The dimensions are L × L × 6 nm3 . The system was relaxed multiple times from
an initial state with random magnetization. The lowest energy states are the leaf state, the C-state,
and the vortex state for L = 80 nm, L = 150 nm, and L = 200 nm, respectively. For each state,
the relative contributions of the exchange energy and the magnetostatic energy to the total energy
are given

for the formation of a closure state is too high. As a compromise the magnetization
bends towards the surface near the edges of the sample. Depending on the size,
the leaf state [51] or the C-state [52] or the vortex state has the lowest energy.
Figure 3 shows the different magnetization patterns that can form in thin film
square elements. The results show that with increasing element size the relative
contribution of the magnetostatic energy, Fdemag /(Fex + Fdemag ) decreases. All
micromagnetic examples in this chapter are simulated using FIDIMAG [43]. Code
snippets are given in the appendix.

Crystal Anisotropy Energy

The magnetic properties of a ferromagnetic crystal are anisotropic. Depending on


the orientation of the magnetic field with respect to the symmetry axes of the
crystal, the M(H ) curve reaches the saturation magnetization, Ms , at low or high
field values. Thus easy directions in which saturation is reached in a low field and
hard directions in which high saturation requires a high field are defined. Figure 4
shows the magnetization curve, measured parallel to the easy and hard direction, of
a uniaxial material with strong crystal anisotropy. The initial state is a two domain
state with the magnetization of the domains parallel to the easy axis. The snapshots
364 L. Exl et al.

Fig. 4 Initial magnetization curves with the field applied in the easy direction (dashed line) and
the hard direction (solid line) computed for a uniaxial hard magnetic material (Nd2 Fe14√ B at room
temperature: K1 = 4.9 MJ/m3 , μ0 Ms = 1.61 T, A = 8 pJ/m, the mesh size is h = 0.86 A/K1 =
1.1 nm). The magnetization component parallel to the field direction is plotted as a function of the
external field. The field is given in units of HK . The sample shape is thin platelet with the easy
axis in the plane of the film. The sample dimensions are 200 × 200 × 10 nm3 . The insets show
snapshots of the magnetization configuration along the curves. The initial state is the two domain
state shown at the lower left of the figure

of the magnetic states show that domain wall motion occurs along the easy axis and
rotation of the magnetization occurs along the hard axis.
The crystal anisotropy energy is the work done by the external field to move
the magnetization away from a direction parallel to the easy axis. The functional
form of the energy term can be obtained phenomenologically. The energy density,
eani (m), is expanded in a power series in terms of the direction cosines of the
magnetization. Crystal symmetry is used to decrease the number of coefficients.
The series is truncated after the first two non-constant terms.

Cubic Anisotropy
Let a, b, and c be the unit vectors along the axes of a cubic crystal. The crystal
anisotropy energy density of a cubic crystal is


eani (m) = K0 + K1 (a · m)2 (b · m)2 + (b · m)2 (c · m)2 + (c · m)2 (a · m)2

+ K2 (a · m)2 (b · m)2 (c · m)2 + . . . .
(57)
7 Micromagnetism 365

The anisotropy constants K0 , K1 , and K2 are functions of temperature. The first


term is independent of m and thus can be dropped since only the change of the
energy with respect to the direction of the magnetization is of interest.

Uniaxial Anisotropy
In hexagonal or tetragonal crystals, the crystal anisotropy energy density is usually
expressed in terms of sin θ , where θ is the angle between the c-axis and the
magnetization. The crystal anisotropy energy of a hexagonal or tetragonal crystal
is

eani (m) = K0 + K1 sin2 (θ ) + K2 sin4 (θ ) + . . . . (58)

In numerical micromagnetics, it is often more convenient to use


eani (m) = −K1 (c · m)2 + . . . . (59)

as expression for a uniaxial crystal anisotropy energy density. Here we used the
identity sin2 (θ ) = 1 − (c · m)2 , dropped two constant terms, namely, K0 and K1 ,
and truncated the series. When keeping only the terms which are quadratic in m,
the crystal anisotropy energy can be discretized as quadratic form involving only a
geometry-dependent matrix.
The crystalline anisotropy energy is

Eani = eani (m)dV , (60)
V

whereby the integral is over the volume, V , of the magnetic body.

Anisotropy Field
An important material parameter, which is commonly used, is the anisotropy
field, HK . The anisotropy field is a fictitious field that mimics the effect of the
crystalline anisotropy. If the magnetization vector rotates out of the easy axis, the
crystalline anisotropy creates a torque that brings M back into the easy direction.
The anisotropy field is parallel to the easy direction, and its magnitude is such that
for deviations from the easy axis, the torque on M is the same as the torque by the
crystalline anisotropy. If the energy depends on the angle, θ , of the magnetization
with respect to an axis, the torque, T , on the magnetization is the derivative of the
energy density, e, with respect to the angle [20]

∂e
T = . (61)
∂θ

Let θ be a small angular deviation of M from the easy direction. The energy density
of the magnetization in the anisotropy field is
366 L. Exl et al.

eK = −μ0 Ms HK cos(θ ) (62)

the associated torque is

TK = μ0 Ms HK sin(θ ) ≈ μ0 Ms HK θ. (63)

For the crystalline anisotropy energy density

eani = K1 sin2 (θ ) (64)

the torque towards the easy axis is

Tani = 2K1 sin(θ ) cos(θ ) = K1 sin(2θ ) ≈ 2K1 θ. (65)

From the definition of the anisotropy field, namely, TK = Tani , we get

2K1
HK = (66)
μ0 Ms

Anisotropy field, easy and hard axis loops: K1 and – depending on the material
to be studied – K2 are input parameters for micromagnetic simulation. The
anisotropy constants can be measured by fitting a calculated magnetization curve to
experimental data. Figure 4 shows the magnetization curves of a uniaxial material
computed by micromagnetic simulations. For simplicity we neglected K2 and
described the crystalline anisotropy with (59). The M(H ) along the hard direction
is almost a straight line until saturation where M(H ) = Ms . Saturation is reached
when H = HK .
The above numerical result can be found theoretically. A field is applied
perpendicular to the easy direction. The torque created by the field tends to increase
the angle, θ , between the magnetization and the easy axis. The torque asserted by
the crystalline anisotropy returns the magnetization towards the easy direction. We
set the total torque to zero to get the equilibrium condition

−μ0 Ms H cos(θ ) + 2K1 sin(θ ) cos(θ ) = 0.

The value of H that makes M parallel to the field is reached when sin(θ ) = 1. This
gives H = 2K1 /(μ0 Ms ). If higher anisotropy constants are taken into account the
field that brings M into the hard axis is H = (2K1 + 4K2 )/(μ0 Ms ).

Magnetoelastic and Magnetostrictive Energy Terms

When the atom positions of a magnet are changed relative to each other the
crystalline anisotropy varies. Owing to magnetoelastic coupling a deformation
produced by an external stress makes certain directions to be energetically more
7 Micromagnetism 367

favorable for the magnetization. Reversely, the magnet will deform in order to
minimize its total free energy when magnetized in certain directions.

Spontaneous Magnetostrictive Deformation


Most generally the spontaneous magnetostrictive deformation is expressed by the
0 as
symmetric tensor strain εij


0
εij = λij kl αk αl , (67)
kl

where λij kl is the tensor of magnetostriction constants. Measurements of the relative


change of length along certain directions owing to saturation of the crystal in
direction α = (α1 , α2 , α3 ) give the magnetostriction constants. For a cubic material,
the following relation holds
 
3 1
εii0 = λ100 αi −
2
, (68)
2 3
3
0
εij = λ111 αi αj for i = j. (69)
2

The magnetostriction constants λ100 and λ111 are defined as follows: λ100 is the
relative change in length measured along [100] owing to saturation of the crystal in
[100]; similarly λ111 is the relative change in length measured along [111] owing
to saturation of the crystal in [111]. The term with 1/3 in (68) results from the
definition of the spontaneous deformation with respect to a demagnetized state with
the averages
αi2 = 1/3 and
αi αj = 0.

Magnetoelastic Coupling Energy


All energy terms discussed in the previous sections can depend on deformations.
The most important change of energy with strain arises from the crystal anisotropy
energy. Thus the crystal anisotropy energy is a function of the magnetization and the
deformation of the lattice. We express the magnetization direction in terms of the
direction cosines of the magnetization α1 = a · m, α2 = b · m, and α3 = c · m (a, b,
and c are the unit lattice vectors) and the deformation in terms of the symmetric
strain tensor εij to obtain

eani = eani (αi , εij ). (70)

A Taylor expansion of (70)

 ∂eani (αi , 0)
eani = eani (αi , 0) + εij (71)
∂εij
ij
368 L. Exl et al.

gives the change of the energy density owing to the strain εij . Owing to symmetry,
the expansion coefficients ∂eani (αi , 0)/∂εij do not dependent on the sign of the
magnetization vector and thus are proportional to αi αj . The second term on the
right-hand side of (71) is the change of the crystal anisotropy energy density
with
 deformation. This term is the magnetoelastic coupling energy density. Using
ij Bij kl αi αj as expansion coefficients, we obtain


eme = Bij kl αi αj εkl , (72)
ij kl

where Bij kl is the tensor of the magnetoelastic coupling constants. For cubic
symmetry, the magnetoelastic coupling energy density is

eme,cubic = B1 (ε11 α12 + ε22 α22 + ε33 α32 )+ 2B2 (ε23 α2 α3 + ε13 α1 α3 + ε12 α1 α2 ) + . . .
(73)

with the magnetoelastic coupling constants B1 = B1111 and B2 = B2323 .


Equation (72) describes change of the energy density owing to the interaction of
magnetization direction and deformation. The magnetoelastic coupling constants
can be derived from the ab initio computation of the crystal anisotropy energy as
function of strain [53]. Experimentally the magnetoelastic coupling constants can
be obtained from the measured magnetostriction constants.
When magnetized in a certain direction, the magnet tends to deform in a way
that minimizes the sum of the magnetoelastic energy density, eme , and of the elastic
energy density of the crystal, eel . The elastic energy density is a quadratic function
of the strain

1 
eel = cij kl εij εkl , (74)
2
ij kl

where cij kl is the elastic stiffness tensor. For cubic crystals the elastic energy is

1
eel,cubic = c1111 (ε11
2
+ ε22
2
+ ε33
2
)+
2
c1122 (ε11 ε22 + ε22 ε33 + ε33 ε11 )+ (75)
2
2c2323 (ε12 + ε23
2
+ ε31
2
).

Minimizing eme + eel with respect to εij under fixed αi gives the equilibrium strain
or spontaneous magnetostrictive deformation

0
εij = εij
0
(Bij kl , cij kl ). (76)
7 Micromagnetism 369

in terms of the magnetoelastic coupling constants and the elastic stiffness constants.
Comparison of the coefficients in (76) and the experimental relation (67) allows to
express the magnetoelastic coupling coefficients in terms of the elastic stiffness con-
stants and the magnetostriction constants. For cubic symmetry the magnetoelastic
coupling constants are

3
B1 = − λ100 (c1111 − c1122 ) (77)
2
B2 = −3λ111 c1212. (78)

External Stress
A mechanical stress of nonmagnetic origin will have an effect on the magnetization
owing to a change of magnetoelastic coupling energy. The magnetoelastic coupling
energy density owing to an external stress σ ext is [54]

eme = − σijext εij
0
. (79)
ij

For cubic symmetry, this gives [20]

3
eme,cubic = − λ100 (σ11 α12 + σ22 α12 + σ33 α22 )
2 (80)
− 3λ111 (σ12 α1 α2 + σ23 α2 α3 + σ31 α3 α1 )

The above results can be derived from the strain induced by the external stress
which is

ext
εij = sij kl σklext , (81)
kl

where sij kl is the compliance tensor. Inserting (81) into (72) gives the magnetoelastic
energy density owing to external stress. For an isotropic material, for example, an
amorphous alloy, we have only a single magnetostriction constant λs = λ100 =
λ111 . For a stress σ along an axis of a unit vector a, the magnetoelastic coupling
energy reduces to

3
eme,isotropic = − λs σ (a · α)2 . (82)
2

This equation has a similar form as that for the uniaxial anisotropy energy density
(59) with an anisotropy constant Kme = 3λs σ/2.

Magnetostrictive Self-Energy
A nonuniform magnetization causes a nonuniform spontaneous deformation owing
to (67). As a consequence, different parts of the magnet do not fit together. To
370 L. Exl et al.

el , will occur. The


compensate this misfit, an additional elastic deformation, εij
associated magnetostrictive self-energy density is

1 
emagstr = el el
cij kl εij εkl . (83)
2
ij kl

el we have to solve an elasticity problem. The total strain,


To compute εij

εij = εij
el
+ εij
0
, (84)

can be derived from a displacement field, u = (u1 , u2 , u3 ), according to [55]


 
1 ∂ui ∂uj
εij = + . (85)
2 ∂xj ∂xi

We start from a hypothetically undeformed, nonmagnetic body. If magnetism is


switched on, εij 0 causes a stress which we treat as virtual body forces. Once these

forces are known, the displacement field can be calculated as usual by linear
elasticity theory. The situation is similar to magnetostatics where the demagnetizing
field is calculated from effective magnetic charges. The procedure is as follows [56].
First we compute the spontaneous magnetostrictive strain for a given magnetization
distribution with (67) or in case of cubic symmetry with (68) and (69). Then we
apply Hooke’s law to compute the stress

σij0 = 0
cij kl εkl (86)
kl

owing to the spontaneous magnetostrictive strain. The stress is interpreted as virtual


body force

 ∂
fi = − σ0. (87)
∂xj ij
j

The forces enter the condition for mechanical equilibrium

 ∂ 
σij = fi with σij = cij kl εkl . (88)
∂xj
j kl

Equations (85) to (88) lead to a system of partial differential equations for the
displacement field u(x). This is an auxiliary problem similar to the magnetostatic
boundary value problem (see section “Magnetostatic Boundary Value Problem”)
which is to be solved for a given magnetization distribution.
7 Micromagnetism 371

Based on the above discussion, we can identify two contributions to the total
magnetic Gibbs free energy: The magnetoelastic coupling energy with an external
stress  
Eme = − σijext εij
0
dV (89)
V ij

and the magnetostrictive self-energy


 
1
Emagstr = cij kl (εij − εij
0
)(εkl − εkl
0
)dV . (90)
2 V ij kl

Artificial multiferroics: The magnetoelastic coupling becomes important in artificial


multiferroic structures where ferromagnetic and piezoelectric elements are com-
bined to achieve a voltage controlled manipulation of the magnetic state [57]. For
example, piezoelectric elements can create a strain on a magnetic tunnel junction
of about 10−3 causing the magnetization to rotate by 90 degrees [58]. Breaking
the symmetry by a stress-induced uniaxial anisotropy, which can be created by a
piezoelectric element, the deterministic switching between two metastable states in
square nano-element is possible as shown in Fig. 5.

 (Ku = 1.32 kJ/m ,


Fig. 5 Simulation of the stress-driven switching of a CoFeB nano-element 3

μ0 Ms = 1.29 T, A = 15 pJ/m, λs = 3 × 10−5 , mesh size h = 0.59 A/(μ0 Ms2 ) = 2 nm,


the magnetostrictive self-energy is neglected). The sample is a thin film element with dimensions
120 × 120 × 2 nm3 . The system switches from 0 to 1 by a compressive stress (−0.164 GPa) and
from 1 to 0 by a tensile stress (0.164 GPa)
372 L. Exl et al.

Characteristic Length Scales

To obtain a qualitative understanding of equilibrium states, it is helpful to consider


the relative weight of the different energy terms towards the total Gibbs free energy.
As shown in Fig. 3, the relative importance of the different energy terms changes
with the size of the magnetic sample. We can see this most easily when we write the
total Gibbs free energy

Etot = Eex + Eext + Edemag + Eani + Eme + Emagstr , (91)

in dimensionless form. From the relative weight of the energy contributions in


dimensionless form, we will derive characteristic length scales which will provide
useful insight into possible magnetization processes depending on the magnet’s size.
Let us assume that Ms is constant over the magnetic body (conditions 2 and 3
in section “Micromagnetics Basics”). We introduce the external and demagnetizing
field in dimensionless form hext = H ext /Ms and hdemag = H demag /Ms and rescale
the length x̃ = x/L, where L is the sample extension. Let us choose L so that
L3 = V . We also normalize the Gibbs free energy E tot = Etot /(μ0 Ms2 V ). The
2
normalization factor, μ0 Ms V , is proportional to the magnetostatic self-energy of
the fully magnetized sample. The energy contributions in dimensionless form are
 2 
lex



ex =
E  x 2 + ∇m
∇m  y 2 + ∇m ,
 z 2 dV (92)
 L2
V

ext
E =− ,
m · hext d V (93)

V

demag = − 1
E ,
m · hdemag d V (94)
2 
V

ani = − K1 ,
E (c · m)2 d V (95)

V μ0 Ms2

where V  is the domain after transformation of the length. Further, we assumed


uniaxial magnetic anisotropy and neglected magnetoelastic coupling and magne-
tostriction. The constant lex in (92) is defined in the following section.

Exchange Length

In (92) we introduced the exchange length



A
lex = . (96)
μ0 Ms2

It describes the relative importance of the exchange energy with respect to the
magnetostatic energy. Inspecting the factor (lex /L)2 in front of the brackets in (92),
7 Micromagnetism 373

we see that the exchange energy contribution increases with decreasing sample
size L. The smaller the sample, the higher is the expense of exchange energy for
nonuniform magnetization. Therefore small samples show a uniform magnetization.
If the magnetization remains parallel during switching, the Stoner-Wohlfarth [59]
model can be applied.
 In the literature, the exchange length is either defined by (96)
 = 2A/(μ M 2 ) [61].
[60] or by lex 0 s

Critical Diameter for Uniform Rotation

In a sphere with uniaxial anisotropy, the magnetization reverses uniformly if its


diameter is below D ≤ Dcrit = 10.2lex [60]. During uniform rotation of the
magnetization, the exchange energy is zero, and the magnetostatic energy remains
constant. It is possible to lower the magnetostatic energy during reversal by
magnetization curling. Then the magnetization becomes nonuniform at the expense
of exchange energy. The total energy will be smaller than for uniform rotation if
the sphere diameter, D, is larger than Dcrit . Nonuniform reversal decreases the
switching field as compared to uniform rotation. The switching fields of a sphere
are [60]

2K1
Hc = for D ≤ Dcrit . (97)
μ0 Ms
2K1 1 34.66A
Hc = − Ms + for D > Dcrit . (98)
μ0 Ms 3 μ0 Ms D 2

In cuboids and particles with polyhedral shape, the nonuniform demagnetizing field
causes a twist of the magnetization near edges or corners [62]. As a consequence
nonuniform reversal occurs for particle sizes smaller than Dcrit . The interplay
between exchange energy and magnetostatic energy also causes a size dependence
of the switching field [63, 64].
Grain size dependence of the coercive field. The coercive field of permanent
magnets decreases with increasing grain size. This can be explained by the different
scaling of the energy terms [64, 65]. The smaller the magnet, the more dominant is
the exchange term. Thus it costs more energy to form a domain wall. To achieve
magnetization reversal, the Zeeman energy of the reversed magnetization in the
nucleus needs to be higher. This can be accomplished by a larger external field.
Figure 6 shows the switching field a Nd2 Fe14 B cube as a function of its edge
length. In addition we give the theoretical switching field for a sphere with the same
volume according to (97) and (98). Magnetization reversal occurs by nucleation
and expansion of reversed domains unless the hard magnetic cube is smaller
than 6lex .
374

Fig. 6 Computed grain size 3


√ dependence of the coercive field of a perfect Nd2 Fe14 B cube at room temperature (K1 = 4.9 MJ/m , μ0 Ms = 1.61 T, A = 8 pJ/m,
the mesh size is h = 0.86 A/K1 = 1.1 nm, the external field is applied at an angle of 10−4 rad with respect to the easy axis). The sample dimensions are
L × L × L nm3 . Left: Switching field as function of L in units of HK . The squares give the switching field of the cube. The dashed line is the theoretical
switching field of a sphere with the same volume. A switching field smaller than HK indicates nonuniform reversal. Right: Snapshots of the magnetic states
during switching for L = 10 nm and L = 80 nm
L. Exl et al.
7 Micromagnetism 375

Wall Parameter

The square root of the ratio of the exchange length and the prefactor of the crystal
anisotropy energy gives another critical length. The Bloch wall parameter

A
δ0 = (99)
K

denotes the relative importance of the exchange energy versus crystalline anisotropy
energy. It determines the width of the transition of the magnetization between two
magnetic domains. In a Bloch wall, the magnetization rotates in a way so that no
magnetic volume charges are created. The mutual competition between exchange
and anisotropy determines the domain wall width: Minimizing the exchange energy
favors wide transition regions, whereas minimizing the crystal anisotropy energy
favors narrow transition regions. In a bulk uniaxial material the wall width is
δB = π δ0 .

Single Domain Size


With increasing particle, the prefactor (lex /L)2 for the exchange energy in (92)
becomes smaller. A large particle can break up into magnetic domains because the
expense of exchange energy is smaller than the gain in magnetostatic energy. In
addition to the exchange energy, the transition of the magnetization in the domain
wall
√ also increases the crystal anisotropy energy. The wall energy per unit area is
4 AK1 . The energy of uniformly magnetized cube is its magnetostatic energy,
Edemag1 = μ0 Ms2 L3 /6. In the two domain states, the magnetostatic energy is
roughly one half
√ of this value, Edemag2 = μ0 Ms2 L3 /12. The energy of the wall
is Ewall2 = 4 AK1 L . Equating the energy of the single domain state, Edemag1 ,
2

with the energy of the two domain state, Edemag2 + Ewall2 , and solving for L give
the single domain size of a cube

48 AK1
LSD ≈ . (100)
μ0 Ms2

The above equation simply means that the energy of a ferromagnetic cube with a
size L > LSD is lower in a the two domain state than in the uniformly magnetized
state. A thermally demagnetized sample with L > LSD most likely will be in a
multidomain state.
We have to keep in mind that the magnetic state of a magnet depends on its
history and whether local or global minima can be accessed over the energy barriers
that separate the different minima. The following situations may arise:

(1) A particle in its thermally demagnetized state is multidomain although L <


LSD [66]. When cooling from the Curie temperature, a particle with L < LSD
may end up in a multidomain state. Although the single domain state has a
lower energy, it cannot be accessed because it is separated from the multidomain
376 L. Exl et al.

state by a high energy barrier. This behavior is observed in small Nd2 Fe14 B
particles [66].
(2) An initially saturated cube with L > LSD will not break up into domains
spontaneously if its anisotropy field is larger than the demagnetizing field.
The sample will remain in an almost uniform state until a reversed domain is
nucleated.
(3) Magnetization reversal of a cube with L < LSD will be nonuniform. Switching
occurs by the nucleation and expansion of a reversed domain for a particle size
down to about 5lex . For example in Nd2 Fe14 B, the single domain limit is LSD ≈
146 nm, and the exchange length is lex = 1.97 nm. The simulation presented in
Fig. 6 shows the transition from uniform to nonuniform reversal which occurs
at L ≈ 6lex .

Mesh Size in Micromagnetic Simulations

The required minimum mesh size in micromagnetic simulations depends on the


process that should be described by the simulations. Here are a few examples:

(1) For computing the switching field of a magnetic particle, we need to describe
the formation of a reversed nucleus. A reversed nucleus is formed near edges
or corners where the demagnetizing field is high. We have to resolve the
rotations of the magnetization that eventually form the reversed nucleus. For
the computation of the nucleation field the required minimum mesh size has to
be smaller than the exchange length [61] at the place where the initial nucleus
is formed.
(2) For the simulation of domain wall motion, the transition of the magnetization
between the domains needs to be resolved. A failure to do so will lead to an
artificial pinning of the domain wall on the computational grid [67]. For the
study of domain wall motion in hard magnetic materials, the required minimum
mesh size has to be smaller than the Bloch wall parameter.
(3) In soft magnetic elements with vanishing crystal or stress-induced ansisotropy,
the magnetization varies continuously [68]. The smooth transitions of the mag-
netization transitions can be resolved with a grid size larger than the exchange
length. Care has to be taken if vortices play a role in the magnetization process
to be studied. Then artificial pinning of vortex cores on the computational grid
[67] has to be avoided.

Brown’s Micromagnetic Equation

In the following, we will derive the equilibrium equations for the magnetization.
The total Gibbs free energy of a magnet is a functional of m(x). To compute an
equilibrium state, we have to find the function m(x) that minimizes Etot taking into
account |m(x)| = 1. In addition the boundary conditions
7 Micromagnetism 377

∇mx · n = 0, ∇my · n = 0, and ∇mz · n = 0 (101)

hold, where n is the surface normal. The boundary conditions follow from (11) and
the respective equations for y and z and applying Green’s first identity to each term
of (14). The boundary conditions (101) can also be understood intuitively [15].
To be in equilibrium, a magnetic moment at the surface has to be parallel with its
neighbor inside when there is no surface anisotropy. Otherwise there is an exchange
torque on the surface spin.
Most problems in micromagnetics can only be solved numerically. Instead
of solving the Euler-Lagrange equation that results from the variation of (91)
numerically, we directly solve the variational problem. Direct methods [69, 70]
represent the unknown function by a set of discrete variables. The minimization
of the energy with respect to these variables gives an approximate solution to the
variational problem. Two well-known techniques are the Euler method and the Ritz
method. Both are used in numerical micromagnetics.

Euler Method: Finite Differences

In finite difference micromagnetics, the solution m(x) is sampled on points


(xi , yj , zk ) so that mij k = m(xi , yj , zk ). On a regular grid with spacing h, the
positions of the grid points are xi = x0 + ih, yj = y0 + j h, and zk = z0 + kh. The
points (xi , yj , zk ) are the cell centers of the computational grid. The magnetization
is assumed to be constant within each cell. To obtain an approximation of the energy
functional, we apply the trapezoidal rule; more precisely, we replace m(x) by the
values at the cell centers mij k and the spatial derivatives of m(x) with the finite
difference quotients. The approximated solution values mij k are the unknowns of an
algebraic minimization problem. The indices i, j , and k run from 1 to the number
of grid points Nx , Ny , Nz in x, y, and z direction, respectively. In the following, we
will derive the equilibrium equations whereby for simplicity we will not take into
account the magnetoelastic coupling energy and the magnetostrictive self-energy.
We can approximate the exchange energy (14) on the finite difference grid as [71]
 
 2Ai+1j k Aij k  mx,i+1j k − mx,ij k 2
Eex ≈ h3 + ··· , (102)
Ai+1j k + Aij k h
ij k

where we introduced the notation Aij k = A(xi , yj , zk ). The prefactor in (102) is


the harmonic mean of the values for the exhange constants in cells i + 1j k and ij k.
This follows from the interface condition Ai+1j k (mx,i+1j k − mx,interface )/(h/2) =
Aij k (mx,interface − mx,ij k )/(h/2), where the minterface is the magnetization at the
interface between the two cells.

Eext ≈ −μ0 h3 Ms,ij k (mij k · H ext,ij k ). (103)
ij k
378 L. Exl et al.

To approximate the magnetostatic energy, we use (42) and (45). Replacing the
integrals with sums over the computational cell, we obtain

μ0   (mij k · n)(mi  j  k  · n )
Edemag ≈ Ms,ij k Ms,i  j  k  |
dSdS  .
8π    ∂Vij k ∂V  
i j k  |x − x
ij k i j k
(104)

The volume integrals in (42) and (45) vanish when we assume that m(x) is constant
within each computational cell ij k. The magnetostatic energy is often expressed in
terms of the demagnetizing tensor Nij k,i  j  k 

μ0 3  
Edemag ≈ h Ms,ij k mTij k Nij k,i  j  k  mi  j  k  Ms,i  j  k  (105)
2   
ij k i j k

We approximate the anisotropy energy (60) by



Eani ≈ h3 eani (mij k ). (106)
ij k

The total energy is now a function of the unknowns mij k . The constraint (5) is
approximated by

|mij k | = 1 (107)

where ij k runs over all computational cells. We obtain the equilibrium equations
from differentiation
⎡ ⎤
∂ ⎣  Lij k
Etot (. . . , mij k , . . . ) + (mij k · mij k − 1)⎦ = 0, (108)
∂mij k 2
ij k
⎡ ⎤
∂ ⎣  Lij k
Etot (. . . , mij k , . . . ) + (mij k · mij k − 1)⎦ = 0. (109)
∂Lij k 2
ij k

In the brackets we added a Lagrange function to take care of the constraints (107).
Lij k are Lagrange multipliers. From (108) we obtain the following set of equations
for the unknowns mij k
 
2Ai+1j k mi+1j k − mij k 2Ai−1j k mi−1j k − mij k
−2Aij k h3 + + · · ·
Ai+1j k + Aij k h2 Ai−1j k + Aij k h2
−μ0 Ms,ij k h3 H ext,ij k (110)
7 Micromagnetism 379


+μ0 Ms,ij k h3 N ij k,i  j  k  mi  j  k  Ms,i  j  k 
i  j  k

∂eani
+h3 = −Lij k mij k .
∂mij k

The term in brackets is the Laplacian discretized on a regular grid. First-order


equilibrium conditions require also zero derivative with respect to the Lagrange
multipliers. This gives back the constraints (107). It is convenient to collect all terms
with the dimensions of A/m to the effective field

H eff,ij k = H ex,ij k + H ext,ij k + H demag,ij k + H ani,ij k. (111)

The exchange field, the magnetostatic field, and the anisotropy field at the compu-
tational cell ij k are

2Aij k 2Ai+1j k mi+1j k − mij k 2Ai−1j k
H ex,ij k = +
μ0 Ms,ij k Ai+1j k + Aij k h2 Ai−1j k + Aij k

mi−1j k − mij k
+ · · · (112)
h2

H demag,ij k = − Nij k,i  j  k  mi  j  k  Ms,i  j  k  (113)
i  j  k

1 ∂eani
H ani,ij k = − , (114)
μ0 Ms,ij k ∂mij k

respectively. The evaluation of the exchange field (112) requires values of mij k
outside the index range [1, Nx ] × [1, Ny ] × [1, Nz ]. These values are obtained by
mirroring the values of the surface cell at the boundary. This method of evaluating
the exchange field takes into account the boundary conditions (101).
Using the effective field, we can rewrite the equilibrium equations

μ0 Ms,ij k h3 H eff,ij k = Lij k mij k . (115)

Equation (115) states that the effective field is parallel to the magnetization at each
computational cell. Instead of (115) we can also write

μ0 Ms,ij k h3 mij k × H eff,ij k = 0. (116)

The expression Ms,ij k h3 mij k is the magnetic moment of computational cell ij k.


Comparison with (1) shows that in equilibrium the torque for each small volume
element h3 (or computational cell) has to be zero. The constraints (107) also have
to be fulfilled in equilibrium.
380 L. Exl et al.

Ritz Method: Finite Elements

Within, the framework of the Ritz method the solution is assumed to depend on a
few adjustable parameters. The minimization of the total Gibbs free energy with
respect to these parameters gives an approximate solution [15, 16]. In the following
we describe a famous and computationally efficient Ritz method, namely, the finite
element ansatz.
Most finite element solvers for micromagnetics use a magnetic scalar potential
for the computation of the magnetostatic energy. This goes back to Brown [16] who

introduced an expression for the magnetostatic energy, Edemag (m, U  ), in terms of
the scalar potential for the computation of equilibrium magnetic states using the

Ritz method. We replace Edemag (m) with Edemag (m, U  ), as introduced in (55), in
the expression for the total energy. The vector m(x) is expanded by means of basis
functions ϕi with local support around node x i

mfe (x) = ϕi (x)mi . (117)
i

Similarly, we expand the magnetic scalar potential



U fe (x) = ϕi (x)Ui . (118)
i

The index i runs over all nodes of the finite element mesh. The expansion
coefficients mi and Ui are the nodal values of the unit magnetization vector and
the magnetic scalar potential, respectively. We assume that the constraint |m| = 1
is fulfilled only at the nodes of the finite element mesh. We introduce a Lagrange
function; Li are the Lagrange multipliers at the nodes of the finite element mesh.
By differentiation with respect to mi , Ui , and Li , we obtain the equilibrium
conditions
 
∂  Li
Etot (. . . , mi , Ui . . . ) + (mi · mi − 1) = 0, (119)
∂mi 2
i

 
∂  Li
Etot (. . . , mi , Ui . . . ) + (mi · mi − 1) = 0, (120)
∂Ui 2
i

 
∂  Li
Etot (. . . , mi , Ui . . . ) + (mi · mi − 1) = 0. (121)
∂Li 2
i

From (119) we obtain the following set of equations for the unknowns mi
7 Micromagnetism 381


2 A∇ϕi · ∇ϕj dV mj
j V

− μ0 Ms H ext ϕi dV
V
 (122)
+ μ0 Ms ∇U ϕi dV
V
 
∂eani ( j ϕj mj )
+ dV = −Li mi .
V ∂mi

Equation (120) is the discretized form of the partial differential equation (50) for the
magnetic scalar potential. Equation (121) gives back the constraint |m| = 1.
In the following, we introduce the effective field at the nodes of the finite element
mesh


2 
H eff,i = − A∇ϕi · ∇ϕj dV mj
μ0 M V
j
 (123)
1 ∂eani
+ H ext,i + H demag,i − dV ,
μ0 M V ∂mi


where M = V Ms ϕi dV . H demag,i is the demagnetizing field averaged over the
finite elements surrounding node i. This average can be computed by plugging (118)
into the third line of (122) and dividing the resulting expression by −μ0 M.
The equilibrium equations are

μ0 MH eff,i = Li mi . (124)

We can write the equilibrium conditions in terms of a cross product of the magnetic
moment, Mmi , and the effective field at node i

μ0 Mmi × H eff,i = 0. (125)

The system is in equilibrium if the torque equals zero and the constraint |mi | = 1 is
fulfilled on all nodes of the finite element mesh.
Instead of a Lagrange function for keeping the constraint |m| = 1, projection
methods [72] are commonly used in fast micromagnetic solvers [73]. In the iterative
scheme for solving (125), the search direction d k+1 i is projected onto a plane
perpendicular to mki , corresponding to first-order approximation of the constraint
at node i. After each iteration k, the vector mk+1
i is normalized.
382 L. Exl et al.

Magnetization Dynamics

Brown’s equations describes the conditions for equilibrium. In many applications,


the response of the system to a time varying external field is important. The
equations by Landau-Lifshitz [74] or Gilbert [75] describes the time evolution of
the magnetization. The Gilbert equation in Landau-Lifshitz form

∂m |γ |μ0 |γ |μ0 α
=− m × H eff − m × (m × H eff ) (126)
∂t 1 + α2 1 + α2

is widely used in numerical micromagnetics. Here |γ | = 1.76086 × 1011 s−1 T−1 is


the gyromagnetic ratio and α is the Gilbert damping constant. In (126) the unit
vector of the magnetization and the effective field at the grid point of a finite
difference grid or finite element mesh may be used for m and H eff . The first term of
(126) describes the precession of the magnetization around the effective field. The
last term of (126) describes the damping. The double cross product gives the motion
of the magnetization towards the effective field.
The interplay between the precession and the damping terms leads to damped
oscillations of the magnetization around its equilibrium state. In the limiting case of
small deviations from equilibrium and uniform magnetization, the amplitude of the
oscillations decay as [76]

a(t) = Ce−t/t0 . (127)

For small damping, the oscillations decay time is [76]

2
t0 = . (128)
αγ μ0 Ms

Switching of magnetic nano-elements. Small thin film nano-elements are key


building blocks of magnetic sensor and storage applications. By application of a
short field pulse, a thin film nano-element can be switched. After reversal, the system
relaxes to its equilibrium state by damped oscillations. Figure 7 shows the switching
dynamics of a NiFe film with a length of 100 nm, a width of 20 nm, and a thickness
of 2 nm. In equilibrium the magnetization is parallel to the long axis of the particle
(x axis). A Gaussian field pulse (dotted line in Fig. 7) is applied in the (-1,-1,-1)
direction. After the field is switched off the magnetization oscillates towards the
long axis of the film. From an exponential fit to the envelope of the magnetization
component, My (t), parallel to the short axes, we derived the characteristic decay
times of the oscillation which are t0 ≈ 0.613 ns and t0 ≈ 0.204 ns for a damping
constant of α = 0.02 and α = 0.06, respectively. According to (128), the difference
between the two relaxations times is a factor of 3, given by the ratio of the damping
constants.
7 Micromagnetism

Fig. 7 Switching of a thin film nano-element by  a short field pulse in the (-1,-1,-1) direction for α = 0.06 (top row) and α = 0.02 (bottom row). (K1 = 0,
μ0 Ms = 1 T, A = 10 pJ/m, mesh size h = 0.56 A/(μ0 Ms2 ) = 2 nm). The sample dimensions are 100 × 20 × 2 nm3 . The sample is originally magnetized in
the +x direction. Left: Magnetization as function of time. The thin dotted line gives the field pulse, Hext (t). Once the field is switched off damped oscillations
occur which are clearly seen in My (t). The bold grey line is a fit to the envelope of the magnetization component parallel to the short axis. Right: Transient
383

magnetic states. The numbers correspond to the black dots in the plot of My (t) on the left
384 L. Exl et al.

Acknowledgments The authors thank the Austrian Science Fund (FWF) under grant No. F4112
SFB ViCoM and grant No. P31140-N32 for financial support. The financial support by the
Austrian Federal Ministry for Digital and Economic Affairs, the National Foundation for Research,
Technology and Development and the Christian Doppler Research Association is gratefully
acknowledged.

Appendix

The intrinsic material properties listed in Table 1 are taken from [77]. The exchange
lengths
√ and the wall parameter are calculated as follows: l ex = A/(μ0 Ms2 ), δ0 =
A/|K1 |.

Table 1 Intrinsic magnetic properties and characteristic lengths of selected magnetic materials
Material TC (K) μ0 Ms (T) A(pJ/m) K1 (kJ/m3 ) lex (nm) δ0 (nm)
Fe 1044 2.15 22 48 2.4 21
Co 1360 1.82 31 410 3.4 8.7
Ni 628 0.61 8 -5 5.2 40
Ni0.8 Fe0.2 843 1.04 10 -1 3.4 100
CoPt 840 1.01 10 4900 3.5 1.4
Nd2 Fe14 B 588 1.61 8 4900 2.0 1.3
SmCo5 1020 1.08 12 17200 3.6 0.8
Sm2 Co17 1190 1.25 16 4200 3.6 2.0
Fe3 O4 860 0.6 7 -13 4.9 23

The examples given in Figs. 3 to 7 were computed using the micromagnetic


simulation environment FIDIMAG [43]. FIDIMAG solves finite difference micro-
magnetic problems using a Python interface. The reader is encouraged to run
computer experiments for further exploration of micromagnetism. In the following
we illustrate the use of the Python interface for simulating the switching dynamics
of a magnetic nano-element (see Fig. 7). The function relax_system computes
the initial magnetic state. The function apply_field computes the response of the
magnetization under the influence of a time varying external field.

import numpy a s np
from f i d i m a g . m i c r o import Sim
from f i d i m a g . common import CuboidMesh
from f i d i m a g . m i c r o import UniformExchange , Demag
from f i d i m a g . m i c r o import TimeZeeman

mu0 = 4 ∗ np . p i ∗ 1 e−7
7 Micromagnetism 385

A = 1 . 0 e −11
Ms = 1 . / mu0

d e f r e l a x _ s y s t e m ( mesh ) :
sim = Sim ( mesh , name= ’ r e l a x ’ )
sim . d r i v e r . s e t _ t o l s ( r t o l =1e −10 , a t o l =1e −10)
sim . d r i v e r . a l p h a = 0 . 5
sim . d r i v e r . gamma = 2 . 2 1 1 e5
sim . Ms = Ms
sim . d o _ p r e c e s s i o n = F a l s e
sim . s e t _ m ( ( 0 . 5 7 7 3 5 0 2 6 9 , 0 . 5 7 7 3 5 0 2 6 9 , 0 . 5 7 7 3 5 0 2 6 9 ) )

sim . add ( U n i f o r m E x c h a n g e (A=A) )


sim . add ( Demag ( ) )

sim . r e l a x ( )
np . s a v e ( ’m0 . npy ’ , sim . s p i n )

d e f a p p l y _ f i e l d ( mesh ) :
sim = Sim ( mesh , name= ’ dyn ’ )
sim . d r i v e r . s e t _ t o l s ( r t o l =1e −10 , a t o l =1e −10)
sim . d r i v e r . a l p h a = 0 . 0 2
sim . d r i v e r . gamma = 2 . 2 1 1 e5
sim . Ms = Ms
sim . s e t _ m ( np . l o a d ( ’m0 . npy ’ ) )

sim . add ( U n i f o r m E x c h a n g e (A=A) )


sim . add ( Demag ( ) )

s i g m a = 0 . 1 e−9
def gaussian_fun ( t ) :
r e t u r n np . exp ( −0.5 ∗ ( ( t −3∗ s i g m a ) / s i g m a ) ∗ ∗ 2 )

mT = 0 . 0 0 1 / mu0
zeeman = TimeZeeman ([ −100 ∗ mT, −100 ∗ mT, −100 ∗
mT] , t i m e _ f u n = g a u s s i a n _ f u n , name= ’H ’ )
sim . add ( zeeman , s a v e _ f i e l d = T r u e )
sim . r e l a x ( d t = 1 . e −12 , m a x _ s t e p s = 10000)

i f __name__ == ’ __main__ ’ :
mesh = CuboidMesh ( nx =50 , ny =10 , nz =1 , dx =2 , dy =2 ,
dz =2 , u n i t _ l e n g t h =1e −9)
r e l a x _ s y s t e m ( mesh )
a p p l y _ f i e l d ( mesh )
386 L. Exl et al.

References
1. Fukuda, H., Nakatani, Y.: Recording density limitation explored by head/media co-
optimization using genetic algorithm and GPU-accelerated LLG. IEEE Trans. Magn. 48(11),
3895–3898 (2012)
2. Greaves, S., Katayama, T., Kanai, Y., Muraoka, H.: The dynamics of microwave-assisted
magnetic recording. IEEE Trans. Magn. 51(4), 1–7 (2015)
3. Wang, H., Katayama, T., Chan, K.S., Kanai, Y., Yuan, Z., Shafidah, S.: Optimal write head
design for perpendicular magnetic recording. IEEE Trans. Magn. 51(11), 1–4 (2015)
4. Kovacs, A., Oezelt, H., Schabes, M.E., Schrefl, T.: Numerical optimization of writer and media
for bit patterned magnetic recording. J. Appl. Phys. 120(1), 013902 (2016)
5. Vogler, C., Abert, C., Bruckner, F., Suess, D., Praetorius, D.: Areal density optimizations for
heat-assisted magnetic recording of high-density media. J. Appl. Phys. 119(22), 223903 (2016)
6. Makarov, A., Sverdlov, V., Osintsev, D., Selberherr, S.: Fast switching in magnetic tunnel
junctions with two pinned layers: micromagnetic modeling, IEEE Trans. Magn. 48(4), 1289–
1292 (2012)
7. Lacoste, B., de Castro, M.M., Devolder, T., Sousa, R., Buda-Prejbeanu, L., Auffret, S., Ebels,
U., Ducruet, C., Prejbeanu, I., Rodmacq, B., Dieny, B., Vila, L.: Modulating spin transfer
torque switching dynamics with two orthogonal spin-polarizers by varying the cell aspect ratio.
Phys. Rev. B 90(22), 224404 (2014)
8. Hou, Z., Silva, A., Leitao, D., Ferreira, R., Cardoso, S., Freitas, P.: Micromagnetic and
magneto-transport simulations of nanodevices based on mgo tunnel junctions for memory and
sensing applications. Phys. B Condens. Matter 435, 163–167 (2014)
9. Leitao, D.C., Silva, A.V., Paz, E., Ferreira, R., Cardoso, S., Freitas, P.P.: Magnetoresistive
nanosensors: controlling magnetism at the nanoscale. Nanotechnology 27(4), 045501 (2015)
10. Ennen, I., Kappe, D., Rempel, T., Glenske, C., Hütten, A.: Giant magnetoresistance: Basic
concepts, microstructure, magnetic interactions and applications. Sensors 16(6), 904 (2016)
11. Sepehri-Amin, H., Ohkubo, T., Nagashima, S., Yano, M., Shoji, T., Kato, A., Schrefl, T.,
Hono, K.: High-coercivity ultrafine-grained anisotropic Nd–Fe–B magnets processed by hot
deformation and the nd–cu grain boundary diffusion process. Acta Mater. 61(17), 6622–6634
(2013)
12. Bance, S., Oezelt, H., Schrefl, T., Winklhofer, M., Hrkac, G., Zimanyi, G., Gutfleisch, O.,
Evans, R., Chantrell, R., Shoji, T., Yano, M., Sakuma, N., Kato, A., Manabe, A.: High energy
product in Battenberg structured magnets. Appl. Phys. Lett. 105(19), 192401 (2014)
13. Fukunaga, H., Hori, R., Nakano, M., Yanai, T., Kato, R., Nakazawa, Y.: Computer simulation
of coercivity improvement due to microstructural refinement. J. Appl. Phys. 117(17), 17A729
(2015)
14. Evans, R.F., Fan, W.J., Chureemart, P., Ostler, T.A., Ellis, M.O., Chantrell, R.W.: Atomistic
spin model simulations of magnetic nanomaterials. J. Phys. Condens. Matter 26(10), 103202
(2014)
15. Brown Jr, W.F.: Micromagnetics, domains, and resonance. J. Appl. Phys. 30(4), S62–S69
(1959)
16. Brown, W.F.: Micromagnetics. Interscience Publishers, New York/London (1963)
17. Heisenberg, W.: Zur Theorie des Ferromagnetismus. Zeitschrift für Physik 49(9), 619–636
(1928)
18. Vleck, V.: The Theory of Electric and Magnetic Susceptibilities. Oxford University Press,
London (1932)
19. Harashima, Y., Terakura, K., Kino, H., Ishibashi, S., Miyake, T.: Nitrogen as the best interstitial
dopant among x = B, C, N, O, and F for strong permanent magnet NdFe11 TiX: First-principles
study. Phys. Rev. B 92, 184426 (2015)
20. Cullity, B.D., Graham, C.D.: Introduction to Magnetic Materials, 2nd edn. IEEE Press,
Hoboken (2009)
21. Kneller, E.: Ferromagnetismus. Springer, Berlin/Göttingen/Heidelberg (1962)
7 Micromagnetism 387

22. Kittel, C.: Introduction to Solid State Physics, 8th edn. Wiley, New York (2005)
23. Talagala, P., Fodor, P., Haddad, D., Naik, R., Wenger, L., Vaishnava, P., Naik, V.: Determination
of magnetic exchange stiffness and surface anisotropy constants in epitaxial Ni1−x Cox (001)
films. Phys. Rev. B 66(14), 144426 (2002)
24. Eyrich, C., Zamani, A., Huttema, W., Arora, M., Harrison, D., Rashidi, F., Broun, D., Heinrich,
B., Mryasov, O., Ahlberg, M., Karis, O., Jönsson, P., From, M., Zhu, X., Girt, E.: Effects of
substitution on the exchange stiffness and magnetization of co films. Phys. Rev. B 90(23),
235408 (2014)
25. Weissmüller, J., McMichael, R.D., Michels, A., Shull, R.: Small-angle neutron scattering by
the magnetic microstructure of nanocrystalline ferromagnets near saturation. J. Res. Natl. Inst.
Stand. Technol. 104(3), 261 (1999)
26. Ono, K., Inami, N., Saito, K., Takeichi, Y., Yano, M., Shoji, T., Manabe, A., Kato, A., Kaneko,
Y., Kawana, D., Yokoo, T., Itoh, S.: Observation of spin-wave dispersion in nd-fe-b magnets
using neutron brillouin scattering. J. Appl. Phys. 115(17), 17A714 (2014)
27. Smith, N., Markham, D., LaTourette, D.: Magnetoresistive measurement of the exchange
constant in varied-thickness permalloy films. J. Appl. Phys. 65(11), 4362–4365 (1989)
28. Livingston, J., McConnell, M.: Domain-wall energy in cobalt-rare-earth compounds. J. Appl.
Phys. 43(11), 4756–4762 (1972)
29. Livingston, J.: Magnetic domains in sintered Fe-Nd-B magnets. J. Appl. Phys. 57(8), 4137–
4139 (1985)
30. Newnham, S., Jakubovics, J., Daykin, A.: Domain structure of thin NdFeB foils. J. Magn.
Magn. Mater. 157, 39–40 (1996)
31. Jackson, J.D.: Classical Electrodynamics, 3rd edn. Wiley, Singapore (1999)
32. Steele, C.: Iterative algorithm for magnetostatic problems with saturable media. IEEE Trans.
Magn. 18(2), 393–396 (1982)
33. Senanan, K., Victora, R.: Effect of medium permeability on the perpendicular recording
process. Appl. Phys. Lett. 81(20), 3822–3824 (2002)
34. Brown Jr, W.F.: Domains, micromagnetics, and beyond: Reminiscences and assessments. J.
Appl. Phys. 49(3), 1937–1942 (1978)
35. LaBonte, A.: Two-dimensional bloch-type domain walls in ferromagnetic films. J. Appl. Phys.
40(6), 2450–2458 (1969)
36. Schabes, M., Aharoni, A.: Magnetostatic interaction fields for a three-dimensional array of
ferromagnetic cubes. IEEE Trans. Magn. 23(6), 3882–3888 (1987)
37. Aharoni, A.: Demagnetizing factors for rectangular ferromagnetic prisms. J. Appl. Phys. 83(6),
3432–3434 (1998)
38. Dittrich, R., Scholz, W.: Calculator for magnetostatic energy and demagnetizing factor. http://
www.magpar.net/static/magpar/doc/html/demagcalc.html (2016) Accessed 11 Aug 2016
39. Sato, M., Ishii, Y.: Simple and approximate expressions of demagnetizing factors of uniformly
magnetized rectangular rod and cylinder. J. Appl. Phys. 66(2), 983–985 (1989)
40. Asselin, P., Thiele, A.: On the field lagrangians in micromagnetics. IEEE Trans. Magn. 22(6),
1876–1880 (1986)
41. Donahue, M., Porter, D., McMichael, R., Eicke, J.: Behavior of μmag standard problem no. 2
in the small particle limit. J. Appl. Phys. 87, 5520–5522 (2000)
42. Vansteenkiste, A., Leliaert, J., Dvornik, M., Helsen, M., Garcia-Sanchez, F., Van Waeyenberge,
B.: The design and verification of MuMax3. AIP Adv. 4(10), 107133 (2014)
43. Bisotti, M.-A., Cortés-Ortuño, D., Pepper, R., Wang, W., Beg, M., Kluyver, T., Fangohr, H.:
Fidimag–a finite difference atomistic and micromagnetic simulation package. J. Open Res.
Softw. 6, 1 (2018)
44. Abert, C., Selke, G., Kruger, B., Drews, A.: A fast finite-difference method for micromagnetics
using the magnetic scalar potential. IEEE Trans. Magn. 48(3), 1105–1109 (2012)
45. Fu, S., Cui, W., Hu, M., Chang, R., Donahue, M.J., Lomakin, V.: Finite-difference micromag-
netic solvers with the object-oriented micromagnetic framework on graphics processing units.
IEEE Trans. Magn. 52(4), 1–9 (2016)
388 L. Exl et al.

46. Scholz, W., Fidler, J., Schrefl, T., Suess, D., Dittrich, R., Forster, H., Tsiantos, V.: Scalable
parallel micromagnetic solvers for magnetic nanostructures. Comput. Mater. Sci. 28(2), 366–
383 (2003)
47. Fischbacher, T., Franchin, M., Bordignon, G., Fangohr, H.: A systematic approach to mul-
tiphysics extensions of finite-element-based micromagnetic simulations: Nmag. IEEE Trans.
Magn. 43(6), 2896–2898 (2007)
48. Abert, C., Exl, L., Bruckner, F., Drews, A., Suess, D.: magnum.fe: A micromagnetic finite-
element simulation code based on FEniCS. J. Magn. Magn. Mater. 345, 29–35 (2013)
49. Chang, R., Li, S., Lubarda, M., Livshitz, B., Lomakin, V.: Fastmag: Fast micromagnetic
simulator for complex magnetic structures. J. Appl. Phys. 109(7), 07D358 (2011)
50. Forster, H., Schrefl, T., Dittrich, R., Scholz, W., Fidler, J.: Fast boundary methods for
magnetostatic interactions in micromagnetics. IEEE Trans. Magn. 39(5), 2513–2515 (2003)
51. Cowburn, R., Welland, M.: Micromagnetics of the single-domain state of square ferromagnetic
nanostructures. Phys. Rev. B 58(14), 9217 (1998)
52. Goll, D., Schütz, G., Kronmüller, H.: Critical thickness for high-remanent single-domain
configurations in square ferromagnetic thin platelets. Phys. Rev. B 67(9), 094414 (2003)
53. James, P., Eriksson, O., Hjortstam, O., Johansson, B., Nordström, L.: Calculated trends of the
magnetostriction coefficient of 3D alloys from first principles. Appl. Phys. Lett. 76(7), 915–917
(2000)
54. Hubert, A., Schäfer, R.: Magnetic Domains, corrected, 3rd printing ed. Springer, Berlin/New
York (2009)
55. Rand, O., Rovenski, V.: Analytical Methods in Anisotropic Elasticity. Birkhäuser,
Boston/Basel/Berlin (2005)
56. Shu, Y., Lin, M., Wu, K.: Micromagnetic modeling of magnetostrictive materials under
intrinsic stress. Mech. Mater. 36(10), 975–997 (2004)
57. Peng, R.-C., Hu, J.-M., Momeni, K., Wang, J.-J., Chen, L.-Q., Nan, C.-W.: Fast 180o
magnetization switching in a strain-mediated multiferroic heterostructure driven by a voltage.
Sci. Rep. 6, 27561 (2016)
58. Zhao, Z., Jamali, M., D’Souza, N., Zhang, D., Bandyopadhyay, S., Atulasimha, J., Wang, J.-P.:
Giant voltage manipulation of mgo-based magnetic tunnel junctions via localized anisotropic
strain: A potential pathway to ultra-energy-efficient memory technology. Appl. Phys. Lett.
109(9), 092403 (2016)
59. Stoner, E.C., Wohlfarth, E.: A mechanism of magnetic hysteresis in heterogeneous alloys.
Philos. Trans. R. Soc. Lond. A Math. Phys. Eng. Sci. 240(826), 599–642 (1948)
60. Skomski, R.: Nanomagnetics. J. Phys. Condens. Matter 15(20), R841 (2003)
61. Rave, W., Ramstöck, K., Hubert, A.: Corners and nucleation in micromagnetics. J. Magn.
Magn. Mater. 183(3), 329–333 (1998)
62. Schabes, M.E., Bertram, H.N.: Magnetization processes in ferromagnetic cubes. J. Appl. Phys.
64(3), 1347–1357 (1988)
63. Schmidts, H., Kronmüller, H.: Size dependence of the nucleation field of rectangular ferromag-
netic parallelepipeds. J. Magn. Magn. Mater. 94(1), 220–234 (1991)
64. Bance, S., Seebacher, B., Schrefl, T., Exl, L., Winklhofer, M., Hrkac, G., Zimanyi, G., Shoji,
T., Yano, M., Sakuma, N., Ito, M., Kato, A., Manabe, A.: Grain-size dependent demagnetizing
factors in permanent magnets. J. Appl. Phys. 116(23), 233903 (2014)
65. Thielsch, J., Suess, D., Schultz, L., Gutfleisch, O.: Dependence of coercivity on length ratios in
sub-micron Nd2 Fe14 B particles with rectangular prism shape. J. Appl. Phys. 114(22), 223909
(2013)
66. Crew, D., Girt, E., Suess, D., Schrefl, T., Krishnan, K., Thomas, G., Guilot, M.: Magnetic
interactions and reversal behavior of Nd2 Fe14 B particles diluted in a Nd matrix. Phys. Rev. B
66(18), 184418 (2002)
67. Donahue, M., McMichael, R.: Exchange energy representations in computational micromag-
netics. Phys. B Condens. Matter 233(4), 272–278 (1997)
68. Schäfer, R.: Domains in extremely soft magnetic materials. J. Magn. Magn. Mater. 215, 652–
663 (2000)
7 Micromagnetism 389

69. Courant, R., Hilbert, D.: Methods of Mathematical Physics. Interscience Publishers, New York
(1953)
70. Komzsik, L.: Applied Calculus of Variations for Engineers. CRC Press, Boca Raton/Lon-
don/New York (2009)
71. Victora, R.: Micromagnetic predictions for magnetization reversal in coni films. J. Appl. Phys.
62(10), 4220–4225 (1987)
72. Cohen, R., Lin, S.-Y., Luskin, M.: Relaxation and gradient methods for molecular orientation
in liquid crystals. Comput. Phys. Commun. 53(1–3), 455–465 (1989)
73. Exl, L., Bance, S., Reichel, F., Schrefl, T., Stimming, H.P., Mauser, N.J.: LaBonte’s method
revisited: An effective steepest descent method for micromagnetic energy minimization. J.
Appl. Phys. 115(17), 17D118 (2014)
74. Landau, L.D., Lifshitz, E.: On the theory of the dispersion of magnetic permeability in
ferromagnetic bodies. Phys. Z. Sowjetunion 8(153), 101–114 (1935)
75. Gilbert, T.: A lagrangian formulation of the gyromagnetic equation of the magnetization field.
Phys. Rev. 100, 1243 (1955)
76. Miltat, J., Albuquerque, G., Thiaville, A.: An introduction to micromagnetics in the dynamic
regime. In: Hillebrands, B., Ounadjela, K. (eds.) Spin Dynamics in Confined Magnetic
Structures, pp. 1–34. Springer, Berlin (2002)
77. Coey, M.D.J.: Magnetism and Magnetic Materials:. Cambridge University Press, Cambridge
001 (2001)

Lukas Exl studied mathematics and computational physics and


received his PhD from TU-Wien in 2014. He is currently running
the project “Reduced Order Approaches in Micromagnetism” at
WPI. He works on computational methods in magnetism and
quantum mechanics with emphasis on (data-driven) PDEs and
model reduction. He is Senior Scientist at the University of
Vienna and lectures numerical methods.

Dieter Suess received his PhD from the TU-Wien in 2002 where
he completed his Habilitation in 2007 in “Computational Material
Science.” In 2006 he proposed “Exchange Spring Media” for
recording. Since 2018 he is assoc. Prof. and Group Speaker of
the “Physics of Functional Materials” group at the University of
Vienna.
390 L. Exl et al.

Thomas Schrefl received his PhD from the TU-Wien in 2002


where he completed his Habilitation in 1999 in “Computational
Physics.” He worked on the development of numerical micromag-
netic solvers for application in magnetic recording and permanent
magnet. He his head of the Center for Modelling and Simulation
at Danube University Krems, Austria
Magnetic Domains
8
Rudolf Schäfer

Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 392
Relevance of Domains and Domain Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 393
Domain Formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 398
Magnetic Energies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 399
Driving Forces for Domain Formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 404
Interplay of Energies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 406
Domain Classification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 409
Domain Walls . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 414
Domain Wall Types . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 415
Domain Wall Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 420
Current-Driven Domain Wall Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 426
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 432

Abstract

Magnetic domains are the basic elements of the magnetic microstructure of


magnetically ordered materials. They are formed to minimize the total energy,
with the stray field energy being the most significant contribution. The reordering
of domains in magnetic fields determines the magnetization curve, domains can
be engineered on purpose, and they can be applied in devices. In this chapter
a review of the basics of magnetic domains is presented. It will be shown how
the magnetic energies act together to determine the domain character and how

R. Schäfer ()
Institute for Metallic Materials, Leibniz Institute for Solid State and Materials Research (IFW)
Dresden, Dresden, Germany
Institute for Materials Science, Dresden University of Technology, Dresden, Germany
e-mail: r.schaefer@ifw-dresden.de

© Springer Nature Switzerland AG 2021 391


J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_8
392 R. Schäfer

domains can be classified. Domain walls and their dynamics, both field- and
current-driven, will be addressed.

Introduction

According to Fig. 1, magnetically ordered materials may be described by five scale-


dependent hierarchic levels [1]: atomic level theory, level (1), explains the origin and
magnitude of magnetic moments, crystal anisotropy, or magnetoelastic interactions,
and it deals with the arrangement of spins on the crystal lattice sites. The theory
works on a microscopic, sub-nanometer scale level. The other extreme level, the
magnetization Curve in level (5), describes the average magnetization of a specimen
as a function of applied magnetic field and may be seen as a macroscopic descriptive
level.
These two extreme levels are interlinked by magnetic microstructure analysis,
levels (2) to (4) in Fig. 1. Here the individual atomic magnetic moments, defined in

M/Ms 1
5. Magnetization Curve
(independent of scale)
H/Ha
1
lysis
na
eA
2. Micromagnetic tur
Analysis uc
str
cro
Mi
(1 - 1000 nm)
c
eti
gn H
Ma

3. Domain Analysis
(1 - 1000 µm)
4. Phase Analysis
1. Atomic Level Theory (> 100 µm)
(< 1 nm)

Fig. 1 Five descriptive levels of magnetically ordered materials, illustrating the link-function
of magnetic microstructure between atomic foundations and technical applications of magnetic
materials. The anisotropy field Ha , used in the magnetization curve, is defined as Ha = 2K1 /μ0 Ms
with K1 and Ms being the first-order cubic anisotropy constant and saturation magnetization,
respectively. Indicated are the sample dimensions for which the five concepts are applicable. The
domain image in level (3) was obtained by magneto-optical Kerr microscopy on the (100) surface
of an Fe3wt%Si-sheet of 0.5 mm thickness, and the M(H )-loop in level (5) was calculated by
magnetic phase theory for the configuration in level (4). (The domain image at level (3) is adapted
by permission from Ref. [1] (c) Springer 1998)
8 Magnetic Domains 393

level (1), are no longer considered. One rather sums them in a certain neighborhood
and represents the local average over a small volume of magnetic moments by a
classical magnetization vector M – the discrete, quantum mechanical, and statistical
properties of the spins and elementary moments are thus ignored. For a constant
temperature this vector has a constant length, the technical saturation magnetization
Ms . Then one can divide the M-vector by the saturation magnetization, leading
to the unit vector of magnetization m(r) = M(r)/Ms with m2 = 1. For this
mesoscopic approach in the description of magnetic materials, it does not matter
whether a material is ferromagnetic or ferrimagnetic, as the latter is also character-
ized by a net magnetization vector. The purpose of magnetic microstructure analysis
is the determination of the vector field m(r) and its response to a magnetic field.
The magnetization vectors are typically arranged as magnetic domains, so Domain
Analysis – level (3) – is in the center of magnetic microstructure analysis. Level
(2), the continuum theory of micromagnetics, deals with the connecting elements
between the domains, the domain walls, and their substructures. Level (4), phase
theory, ignores the specific arrangement of the domains and rather focuses on their
volume distribution by collecting all domains that are magnetized along a specific
direction in a domain phase. The rearrangement of the phases in a magnetic field
finally leads to the (idealized) magnetization curve. So the domains are at the end
responsible for magnetization curves.
In a conventional definition, magnetic domains are uniformly magnetized regions
that appear spontaneously in otherwise unstructured ferro- or ferrimagnetic samples
[1]. In the example presented in the center of Fig. 1, the domains are well-ordered
owing to the facts that the surface of the crystal is “well”-oriented, meaning
that it contains two easy anisotropy axes for the magnetization, that the crystal
is largely free from internal mechanical stress, and that no magnetic field is
applied to the specimen. In general, however, magnetic domains do not have to be
uniformly magnetized, and they can be more or less complex depending on many
circumstances. To give an impression of the variability of domain patterns, a collage
of selected domain images of various magnetic materials is shown in Fig. 2.
In this chapter a review of some basic aspects of magnetic domains is presented
that is based on an earlier textbook on magnetic domains [1]. The magneto-optical
Kerr images were partly taken and adapted from the book.

Relevance of Domains and Domain Analysis

The magnetization curve in Fig. 1 is calculated under the assumption of an infinitely


extended specimen, the arrangement of domain phases rather than individual
domains is considered, and the properties of domain walls are completely ignored.
In this phase-theoretical approach, it is further assumed that the domain phase
volumes can freely reach their optimum equilibrium values, thus ignoring coer-
civity and irreversibility effects. The result is a reversible, vectorial magnetization
394 R. Schäfer

NiFe film NiFe


film
dCo
Co/Si/G

Co crysta Fe
l
cglass film
talli
me

et
she FeS
Si i sh
eet
Fe Ga
rne
t fi
lm

t
shee
NiFe FeSi

Fig. 2 Collection of domain images, obtained by Kerr microscopy on various magnetic materials.
(Most domain images are adapted by permission from Ref. [1] (c) Springer 1998)

curve that may be used as a theoretical reference. In practice, however, those


conditions are rarely met. Applied materials have a polycrystalline-, amorphous-,
or nanocrystalline microstructure; they are of finite size in the shape of particles,
films, ribbons, sheets, or bulk magnets; domain walls may be pinned at defects
and grain boundaries; or they may interact with each other in case of thin films;
mechanical stress can influence the local preference of the magnetization direction;
and the domains and effective magnetic field are strongly influenced by surfaces and
the sample geometry. All these features may have strong effects on the measured
magnetization curve leading to coercivity and irreversibility. Therefore the correct
interpretation of measured hysteresis curves often requires the experimental analysis
of the domains that are responsible for the loop and of their (in general) irreversible
response to magnetic fields, known as magnetization process.
Figure 3 demonstrates such processes for four different magnetic films with
strong perpendicular anisotropy, i.e., they all have an easy axis for the magnetization
that is aligned perpendicular to the film plane. In the course of the experiments, a
strong positive magnetic field was applied along the anisotropy axis, and then the
field was inverted and successively increased in the opposite direction, thus follow-
ing the upper branch of the shown hysteresis loops. The magnetization process is
initiated by domain nucleation, followed by domain wall motion (upper three rows
8 Magnetic Domains 395

1 -15.5 mT -16.2 mT -17.5 mT -17.9 mT


M/Ms

0
µ0H
in mT
-1 20 µm
-40 -20 0 20 40
1 -7.0 mT -7.8 mT -7.9 mT -8.1 mT

-1 100 µm
-40 -20 0 20 40
1 -200 mT -300 mT -370 mT -400 mT

-1 10 µm
-1000 -500 0 500 1000
1 -1 mT -20 mT -67 mT -106 mT

-1 10 µm
-1000 -500 0 500 1000

Fig. 3 Domain nucleation and growth in magnetic films with strong perpendicular anisotropy,
together with the hysteresis loops along which the domains were imaged. First row: [Co
(0.3 nm)/Pt (0.7 nm)]3 multilayer (sample courtesy D. Makarov, Dresden). Second row: Pt
(3 nm)/Co (1 nm)/Pt (1.5 nm) trilayer (sample courtesy P.M. Shepley and T.A. Moore, Leeds).
Third row: FePt film, 16 nm thick (sample courtesy P. He and S.M. Zhou, Tongji [4]). Fourth
row: FePd(11 nm)/FePt (24 nm) double layer (Sample courtesy L. Ma and S.M. Zhou, Tongji [3])

in the figure) or proceeding nucleation (lower row). For such polycrystalline thin
films, the domain character is more determined by domain wall coercivity effects
due to defects and roughness, randomness of the sample nanostructure, etc. rather
than by an equilibrium of magnetic energies [2, 3]. Typical for such films is a slow
creeping of the nucleated reversed domains into the still saturated area, indicating
thermally activated processes. The domains shown can therefore hardly be predicted
by domain theory – the only way to interpret the hysteresis curves is by observation
of the domains that are responsible for the loops.
This is also true for the M(H )-loop of the amorphous ribbon in Fig. 4. Rather
than revealing a rectangular loop as expected for such material, a predominantly
flat curve is measured inductively with two distinct steps at small field. Domain
observation immediately discloses the reason for such behavior: perpendicularly
magnetized domains across most of the ribbon’s cross section are seen from the side,
whereas longitudinal 180◦ domains are found on the surface. The perpendicular
domains are magnetized by reversible rotational processes in an applied magnetic
396 R. Schäfer

1 Surface M(H)-

ld ic
fie net
curve
M/Ms

ag
(by MOKE

M
magnetometry)

Volume M(H)-
curve (inductively

µm
–1 measured)

20
–8 –6 –4 –2 0 2 4 6 8
H in kA/m

Fig. 4 Hysteresis curves and domains of a surface-crystallized, 20-μm-thick amorphous ribbon


of composition Fe84.3 Cu0.7 Si4 B8 P3 . Together with E. Lopatina, IFW Dresden (Adapted and
reprinted from Ref. [5] with permission from Elsevier)

field along the ribbon axis, leading to the flat portions of the loop, while the surfaces
are magnetized by the coercive motion of 180◦ domain walls, which is responsible
for the low-field behavior. In fact, a rectangular surface loop is measured by
magneto-optical Kerr (MOKE) magnetometry with the same coercivities as found
inductively. From the relative induction amplitude at the switching fields of the
surface, a depth of about one micrometer for the longitudinally magnetized surface
regions can be derived. The reason for this behavior is a crystallized surface that
sets the volume under planar compression stress, whereas the surface itself is under
tension. By magnetoelastic interaction, this inhomogeneous stress state leads to a
perpendicular anisotropy in most of the volume and longitudinal anisotropy at the
surfaces.
There are also cases where the interpretation of a hysteresis loop is apparently
simple: the rectangular loop measured on a permalloy film that is in direct exchange
contact with an antiferromagnetic NiO film (Fig. 5) suggests 180◦ wall motion as
dominating magnetization process. In fact this can be proven by domain imaging.
However, the domain walls surprisingly change character between symmetric Néel
and cross-tie wall (see section “Domain Walls”) on the descending and ascending
branches of the loop, respectively. This subtle difference can by no means be derived
from the magnetization loop. These reversible wall alterations indicate the existence
of bi-modal coupling strengths due to pinned and unpinned spins at the interface
between ferro- and antiferromagnet [6].
Surprising domain phenomena may also be found when magnetic samples are
excited by high-frequency magnetic fields. The ground domain state of a nanocrys-
talline FeNbSiCuB tape wound core with a week circumferential anisotropy, for
example, consists of 180◦ domains that are running along the circumferential
direction (Fig. 6a). Excited quasistatically along the easy axis, i.e., in a slowly
changing magnetic field well below the Hertz regime, 180◦ walls are nucleated,
8 Magnetic Domains 397

M/Ms
10 µm
0
Heb

–1
–40 –20 0 20 40
µ0H in mT

Fig. 5 Field-shifted hysteresis loop in a Ni81 Fe7 (30 nm) / NiO (30 nm) exchange-bias double film
together with high-resolution domain wall images along the forward and backward branch of the
loop. A symmetric Néel wall (upper image) and an asymmetrically distorted cross-tie wall (lower
image) are observed. Obtained at IFW Dresden together with J. McCord. (Adapted and reproduced
from Ref. [6] by permission from IOP Publishing)

At 0.04 Hz At 1 kHz
Ground state
H
1
M/Ms

Ku

1 H in A/m
M/Ms

–10 –5 5 10

H in A/m –1
–5 5

0.2 mm
a) b)

Fig. 6 Difference between quasistatic and dynamic magnetization process, demonstrated for
a nanocrystalline Fe73 Cu1 Nb3 Si16 B7 tape wound core with circumferential anisotropy. The
quasistatic hysteresis loop in (a) is governed by the motion of 180◦ walls along the easy axis,
whereas dynamic excitation with a sinusoidal ac field of 1 kHz frequency results in a domain
nucleation-dominated reversal (b). Together with S. Flohrer, IFW Dresden. (Adapted and reprinted
from Ref. [7] with permission from Elsevier)

shifted and annihilated, leading to a square-type hysteresis loop. If excited at high


frequency, however, the domain character surprisingly changes to a patch-type
pattern (b) and the area of the hysteresis loop, and thus the energy loss increases
due to eddy current effects.
398 R. Schäfer

Without domain refinement After laser scribing


2 mm
2 µm

a) b)

Fig. 7 Engineering and application of domains: (a) Domain refinement in a grain-oriented


transformer steel sheet by laser scribing for the purpose of energy loss reduction. (b) Current-
induced motion of domain walls in a CoNi multilayer microwire with perpendicular anisotropy, an
experimental test structure for the race-track memory [8]. (Courtesy S.S.P. Parkin, IBM and Halle)

From these examples we see that the correct interpretation of hysteresis curves
may indeed be strongly supported by the analysis of the domains that are responsible
for the loop. But magnetic domains can also be engineered on purpose to obtain
favorable properties in devices, and they can be actively used in devices as
functional units. The best-known example for domain engineering is the intended
domain refinement in transformer steel sheets by scratching or laser scribing
(Fig. 7a). The stress introduced locally in this way interrupts the basic domains,
acting like an artificial grain boundary. This mechanism works for well-textured
material, in which the basic domains would otherwise be wide thus causing large
anomalous eddy current losses when operated at power frequency.
A classical example for the application of magnetic domains is the bubble
memory which was developed back in the 1970s. In this device, cylindrical domains
as carriers of information were shifted in a magnetic garnet film along deposited
ferromagnetic structures on top of the film by applying a rotating magnetic field.
A modern concept of such a domain shift register device is the race track memory
(Fig. 7b), in which perpendicularly magnetized domains are shifted in a magnetic
nanowire by electrical current.

Domain Formation

Magnetic domains are formed to minimize the total free energy [9]. Under ideal
conditions (i.e., ignoring coercivity), a vector field of magnetization directions m(r)
arises in a ferro- or ferrimagnetic specimen so that the total energy reaches an
absolute or relative minimum under the constraint of a constant magnetization, i.e.,
m2 = 1. In this section the energies are reviewed, and it is demonstrated how they
act together to define the domain character.
8 Magnetic Domains 399

Magnetic Energies

The relevant magnetic energies are summarized in an illustrative way in Fig. 8. They
can be classified in local terms, which depend on the local magnetization direction
(anisotropy, applied field, and magnetoelastic coupling energy) and nonlocal terms
that give rise to torques on the magnetization vector, which depend at any point on
the magnetization direction at every other point. The stray field and magnetostrictive
self-energies belong to this class. The exchange energy may be seen as local, but it
depends on the derivatives of the magnetization direction. In the following the most
important aspects of the energy terms are listed.  Given are energy densities Ex ,
which by integration lead to the energies εx = Ex dV with V being the sample
volume.

Exchange energy The alignment of magnetic moments occurs via exchange


coupling. Deviations from a constant magnetization direction therefore invoke the
penalty of exchange energy

Exchange energy Anisotropy energy External field energy

E ex = 0 M M
H ext = 0
M Easy
axis Ea = 0

E ex > 0 H ext
M

Ea > 0 H ext

Stray field energy Magnetostrictive Magneto-elastic


divH d m self energy coupling energy
N [100] [100]
S N
S M E ms = 0
N N

Hd 100 >0
0]
M [10 0]
S N [01 0] Tensile
[10
stress
E ms > 0

Fig. 8 Summary of magnetic energies that are relevant for the formation and character of magnetic
domains
400 R. Schäfer

Eex = A(grad m)2 , (1)

where A is the exchange stiffness constant, a temperature-dependent material


constant in units J/m.

Anisotropy energy Most magnetic materials are anisotropic, i.e., there are easy
axes along which the magnetization vector is preferably aligned and along which
the saturated state can be “more easily” obtained than along other directions. These
can be preferred crystal axes (magnetocrystalline anisotropy) or an axis that arises
from some short-range ordering of atoms like Ni-Ni and Fe-Fe atomic pairs in NiFe
alloys. The driving force for this induced anisotropy is the magnetization of the
material that is present at an annealing temperature below the Curie temperature
and which can intentionally be aligned by an applied external magnetic field.
Also annealing under mechanical stress may result in a uniaxial anisotropy, called
creep-induced anisotropy. In amorphous and nanocrystalline ribbons, this type of
anisotropy may be dominating, and often an easy plane of magnetization transverse
to the stress axis is created by stress annealing. Shape effects are part of the stray
field energy, and they do not belong to the anisotropy terms.
Deflecting the magnetization out of an easy axis requires additional energy, called
anisotropy energy. In the most simple case of a uniaxial anisotropy (as it occurs
in crystal lattices with hexagonal or tetragonal symmetry or in case of an induced
anisotropy), the energy density is written as

Ea = Ku1 sin2 ϑ + Ku2 sin4 ϑ , (2)

where ϑ is the angle between anisotropy axis and magnetization direction and Ku1
and Ku2 are the anisotropy constants of first and second order – higher orders can
usually be neglected. An easy axis is described by a large positive Ku1 , whereas
‘planar’ and ‘conical’ anisotropies are found for large negative Ku1 and intermediate
Ku2 values. The anisotropy constant Ku1 corresponds to the energy needed to
saturate the sample in the so-called “hard” direction (ϑ = 90◦ ). In multiaxial
materials such as iron, all 100 directions are easy, whereas in nickel the 111
axes are the preferred crystal axes.
External field energy, also called Zeeman energy, is added to a magnet if an
external magnetic field H ext is applied. It is given by

EZ = −μ0 H ext · M = −μ0 Hext · M · cos(ϕ) , (3)

where ϕ is the angle between magnetization and field. This interaction energy of
external field and magnetization vector field m(r) causes domain wall motion and
rotational processes and finally leads to saturation along the field direction if the
field is strong enough. The minimum of the Zeeman energy is achieved when the
magnetization is aligned to the magnetic field (ϕ = 0).
8 Magnetic Domains 401

Stray field energy Sinks and sources of the magnetization vector field (div m)
lead to magnetic poles, which act as sources and sinks for a magnetic stray field.
Magnetic poles can be present as volume or as surface poles if the magnetization
vector M is not parallel to the surface. The stray field H d , arising from the poles, is
illustrated in Fig. 8 for a finite ellipsoidal magnet that is homogeneously magnetized
to the right. This leads to north (N) and south (S) poles at the edges and a stray field
from (N) to (S). Within the magnet the stray field is called demagnetizing field as it
opposes the magnetization. The presence of poles and stray fields causes the stray
field energy

1
Ed = − μ0 H d · M with H d = −N · M . (4)
2

The demagnetizing factor N (a tensor in general) is zero for infinitely extended


bodies and becomes the larger the closer the specimen edges along M. The stray
field energy thus scales with N and with the average magnetization. A particularly
unfavorable case is an infinitely extended plate that is magnetically saturated
perpendicular to its surface. The demagnetizing factor along M is 1 then, and the
demagnetizing or stray field energy is written

1
Ed = μ0 Ms2 = Kd . (5)
2

The stray field energy coefficient Kd is a measure for the maximum energy densities
which may be connected with stray fields. Independent of the complexity of real
stray fields, their energy always scales with the material parameter Kd . As the
demagnetizing field of a body along a short axis is stronger than along a long axis,
the applied magnetic field along the short axis has to be stronger to produce the same
field inside the specimen. The shape of the magnet is thus the source of magnetic
anisotropy (shape anisotropy). For an infinitely extended body that is magnetized
along an infinite direction, the demagnetizing factor N is zero, and there will be no
stray field energy at all.

Magnetostrictive self-energy In magnetostrictive material, the crystal lattice is


spontaneously elongated or contracted along the magnetization direction if the mag-
netostriction constant λ is positive or negative, respectively. As magnetostriction
is quadratic in the magnetization vector, this lattice distortion is not important for
180◦ domains as all domains will lead to the same deformation. However, for
the 90◦ domain configuration shown in Fig. 8, it will cause elastic energy (called
magnetostrictive self-energy) as the spontaneous deformations of various parts of
the domain pattern, indicated by ellipses in the figure, do not fit together elastically.
The energy density of this incompatible domain configuration is

9
Ems = − Cλ2100 , (6)
8
402 R. Schäfer

λs > 0

λs = +35·10–6 λs = +24·10–6

λs = 0

100 µm
λs = +8·10 –6
λs < 0.2·10–6
a) b)
Fig. 9 Illustration of magnetostrictive and magnetoelastic energies: (a) Closure domains,
observed on the surfaces of two amorphous ribbons with positive and zero magnetostriction as
indicated. A backward pointing magnetic field along the ribbon axis was applied. (b) Typical
domains in the as-quenched state of amorphous ribbons with different magnetostriction constants.
Frozen-in internal mechanical stress, which is present in all four materials after quenching, leads to
different complexity in domains depending on the magnetostriction constant. The ribbon thickness
is about 20 μm in each case

where λ100 is the magnetostriction constant for the 100 directions and C is the
relevant shear modulus for the given configuration. If not enforced for topological
reasons or by magnetic fields, nature tries to avoid such incompatible domain
arrangements. Figure 9a demonstrates this effect for amorphous ribbons with some
induced anisotropy perpendicular to the ribbon surface. A moderate magnetic field
along the ribbon axes was applied that enforces a certain longitudinal magnetization.
If magnetostriction is positive, domains running transverse to the field direction are
more favorable because here the sample elongation in the neighboring domains fits
together elastically. For a material without magnetostriction, a longitudinal domain
arrangement will cause no problem as the elastic distortion, indicated in Fig. 8, will
not occur. In fact this arrangement is even more favorable because here the specific
wall energy of the basic domain walls is lower than for the transverse case.

Magnetoelastic interaction energy There is also an inverse effect: applied


mechanical stress of nonmagnetic origin can act on the magnetization direction
in materials with non-zero magnetostriction by adding magnetoelastic interaction
energy. The stress can be an external stress or some nonmagnetic internal stress
resulting from dislocations or inhomogeneities in composition, structure, and
temperature. In the example shown in Fig. 8, a horizontal tensile stress favors
the horizontal anisotropy axes in a material with otherwise dominating positive
cubic crystal anisotropy. If magnetocrystalline anisotropy is low or absent like in
8 Magnetic Domains 403

amorphous materials, the magnetoelastic coupling may lead to dominating stress-


induced anisotropies (Fig. 9b). This can be seen by writing the magnetoelastic
coupling energy as

3
Eme = λs σ sin2 ϑ (7)
2

where σ is the uniaxial mechanical stress, λs the isotropic magnetostriction constant,


and ϑ the angle between magnetization vector and stress axis. This energy term
describes a uniaxial anisotropy along the stress axis with an anisotropy constant of
Ku = 32 λs σ , compare Eq. (2). Although magnetostriction is a relatively weak effect
with induced strains of typically 10−5 only, the examples in Fig. 9 demonstrate that
its effect on domain patterns can be significant.

Domain wall energy The specific energy of domain walls is not an independent
term, but rather consists of exchange and anisotropy energy, owing to the deviation
of magnetization from the anisotropy axes and non-parallel magnetic moments
across the wall. In the most simple case of a Bloch√wall in an infinitely extended
uniaxial material the specific wall energy is γw0 = 4 A/Ku . See section “Domain
Walls” for more information on domain walls.
For summary, the characteristic coefficients of the magnetic energy terms and
their order of magnitude in typical magnetic materials are collected in Fig. 10.

Magnetic Energy Energy Coefficient Range

Exchange stiffness constant A


Exchange energy 10 J/m
Material constant
Anisotropy constant K
Material constant
Anisotropy energy ±(102 7
) J/m3
K1: Crystal anisotropy
Ku: Induced (uniaxial) anisotropy
µ0HextMs Depends on field
External field energy Hext = external field magnitude Hext,
Ms = saturation magnetization unit J/m3
2
Stray field energy Kd = 1/2 µ0Ms 6
J/m3
λ Depends on stress
Magnetoelastic magnitude ext,
= mechanical stress
interaction energy
λ = magnetostriction constant unit J/m3
2

Magnetostrictive self
C = shear modulus 3
J/m3
energy
λ = magnetostriction constant

Fig. 10 Summary of energy coefficients and their range of magnitude


404 R. Schäfer

Driving Forces for Domain Formation

Primarily it is the stray field energy that is responsible for the development of
magnetic domains: domains are formed to reduce or avoid stray field energy. This
fact is illustrated in Fig. 11 by comparing an infinitely extended with a finite sample.
The NiFe (permalloy) film was sputter-deposited in the presence of a magnetic
field, so it has a weak uniaxial anisotropy along the vertical direction in the images.
For Fig. 11a, a piece of the film was broken from the wafer which extends 30 mm
along the easy axis. In view of the thickness of just 240 nm, the specimen may be
considered as infinitely extended so that stray field energy does not play a role. In
fact, domains are not visible in that case: when a magnetic field along the easy axis is
applied, the film switches from magnetization up to down and vice versa, leading to
a magnetization curve with two steep, discontinuous steps at the coercivity field. The
switching occurs by the fast and abrupt motion of a 180◦ domain wall as indicated
schematically in the figure. The multidomain states at the discontinuity fields cannot
be captured in the experiment because the expense of domain wall energy makes
them statically unfavorable.
The situation changes if an open sample, prone to a demagnetizing field, is
considered. For Fig. 11b the infinite permalloy film was replaced by a 100×100 μm2
film element. For this finite specimen, the in-plane demagnetizing factor N has
raised to 0.0015, and a demagnetizing energy of N Kd m2 is added. A saturated state
at remanence, like in Fig. 11a, would thus be highly unfavorable as it would cost

30 x 5 mm2

M/Ms

1 1
Finite sample
( = 0)
Hext = Hint + M

µ0Hext in mT µ0Hext in mT
–5 5 –5 5

–1 –1
a) b)

Fig. 11 Comparison between infinitely extended and finite samples. Shown are the magneto-
optically measured magnetization curves along the (vertical) induced anisotropy axis in a Ni80 Fe20
Permalloy film of 240 nm thickness, which is infinitely extend in (a) and of finite size in (b). The
domain images in (b) show the full patterned element, whereas in (a, upper inset) only a part of
the extended film is shown. The lower inset in (a) is a schematics
8 Magnetic Domains 405

maximum stray field energy (| m |= 1). At zero field, the film element is rather
in a demagnetized, multidomain state that reveals a flux-closed, pole-free nature.
In the magnetization curve, this is expressed by the shearing transformation: for
a given magnetization value m, the external field Hext must be enlarged by the
demagnetizing field −Hd = N Ms m to reach the same magnetization state. Thus
the discontinuous magnetization curve is transformed into a finite-slope curve with
a well-defined magnetization value for every field value. For the finite element, the
domains can thus be followed along the sheared M(H )-loop.
So for the existence of magnetic domains, a finite sample size along the
magnetization direction (easy axis) is required as the stray field energy is the driving
force for domain observation. Infinitely extended films will consequently also have
domains if the easy axis is perpendicular to the film plane, compare Fig. 3.
There are two further cases in which magnetic domains or related objects can
exist even in the absence of demagnetizing effects:

• Consider a sample that is embedded in an ideal soft magnetic yoke to obtain flux-
closure. With a coil, wrapped around the yoke, a certain average magnetization
can be enforced in the yoke by some feedback mechanism. If a magnetization
value is then enforced in the sample that lies within the range of a discontinuous
jump in the magnetization curve, a multidomain state will be enforced in which
the two states at the endpoints of the jump will be mixed in a certain volume
ratio so that the enforced magnetization is achieved. In case of the previously
discussed uniaxial material, this would be a 180◦ domain state in which the 180◦
domain walls move as the enforced magnetization is varied. Such circuits are
realized in machines with an inductive load on a rigid voltage, such as an idling
transformer.
• The second case may be found in magnetic materials with broken inversion
symmetry in the atomic lattice in which the crystallographic handedness induces
a quantum-mechanical Dzyaloshinskii-Moriya interaction (DMI) [10] by spin-
orbit scattering. Unlike direct Heisenberg or superexchange, which favor parallel
or antiparallel alignment of neighboring magnetic moments according to a
Hamiltonian that is proportional to S i · S j , the DMI is proportional to S i × S j
thus favoring perpendicularly aligned neighboring spins S. In competition with
collinear coupling, the DMI can lead to nanoscale, homochiral magnetization
modulations like long-period helical spin-spiral phases. Most prominent is the
topologically stable skyrmion spin structure that was predicted theoretically by
A. Bogdanov [11, 12] and which has been directly observed in nanolayers of
cubic helimagnets with intrinsic DMI [13] and in Fe/Ir bilayers [14] with surface-
/interface-induced chiral interactions [15]. Magnetic skyrmions are axisymmetric
vortex patterns with a homochiral rotation of spins that can exist as isolated
entities in the saturated states of chiral magnets [14,16] or in form of skyrmionic
condensates (two-dimensional lattices and other mesophases) [12, 17].

In modern literature, intrinsically caused magnetic modulations (e.g., chiral


helicoids and skyrmions) are often classified as magnetic micro- or spintextures,
406 R. Schäfer

whereas the modulated elements of multidomain states (domain walls, Bloch lines,
Bloch points, magnetic swirls etc. – compare Fig. 18) are commonly addressed as
being part of magnetic microstructure [1]. This different classification, however, is
questionable: spatially inhomogeneous spin structures arising in magnetic nanolay-
ers are formed under the mutual influence of intrinsic and dipolar forces [18]. This
levels out the difference between the terms magnetic microstructure and magnetic
micro- or spintextures [19].

Interplay of Energies

Once the precondition for domain formation is fulfilled, the domain character is
finally determined by an interplay of the magnetic energies. For demonstration of
this principle, let us have a look at the prominent example of domain formation in
grain-oriented Fe3wt%Si steel that is used as core material in transformers. Like
for pure iron, the easy directions of magnetization are the 100 directions for this
material. Transformer sheets are typically 0.3 mm thick, consist of wide grains in the
centimeter regime, and are Goss-textured. In this [001](110) texture the [001] easy
direction is oriented, within a few degrees deviation, along the rolling axis during the
manufacturing process of the sheets. The grain surfaces are (110) oriented within the
same accuracy, and the other two easy axes are oriented at angles of ±45◦ relative
to the surface. As shown in Fig. 12, the domain structure of such sheets consists
of simple slab domains that are separated by 180◦ walls in case of ideally oriented
grains. Their existence may, e.g., be enforced by some demagnetization effects at
the grain boundaries. For increasing out-of-plane misorientation of the [001] easy
direction, fine lancet-shaped domains of increasing density are superimposed on the
basic domains.
The formation of those so-called supplementary domains is a consequence of
energy optimization. Let us firstly assume an infinitely extended grain with ideal
(110) orientation. It will be homogeneously magnetized along the surface-parallel
easy axis (Fig. 13a), thus completely avoiding magnetic poles. As the grain is
infinitely extended, domains are not to be expected. A different situation arises if
the [001]-axis is misoriented by some degrees relative to the surface (b). Assuming
that the magnetization strictly follows the [001] axis, magnetic surface poles will
arise. The associated stray field energy can be reduced by forming ±180◦ basis
domains (c) which leads to the presence of opposite poles on the same surface, thus
allowing the field lines of the stray field to run along the surface. A further reduction
of stray field energy could be achieved by reducing the basic domain spacing as
this would bring the opposite poles in closer distance. However, the narrower the
domain width, the higher the expense of domain wall energy associated with the
rising wall area of the basic domain walls that extend all through the thickness.
Nature finds a more economic way to keep the overall energy low by adding
supplementary domains to the basic domains (d). The shallow lancet domains at
the surface collect the net flux that is transported toward the surface in the basic
domain. The lancets are oppositely magnetized to the basic domains, thus leading to
8 Magnetic Domains 407

(110) surface
Goss texture 100 easy axes

0.1 mm Out-of-plane
misorientation

2° miso
riented

4° miso
riented

idealy
oriente
d

8° mis
oriente
30 mm d

Fig. 12 Domains on a Goss-textured transformer sheet. Shown are the domains of four grains with
increasing out-of-plane misorientation as indicated. The ceramic insulation coating, by which such
sheets are usually covered to avoid eddy currents between the sheets, was removed for domain
imaging by Kerr microscopy. (The domain images are adapted by permission from Ref. [1] (c)
Springer 1998)

a narrow spacing of opposite surface poles as required for stray field reduction. The
flux is then transported to a surface of opposite polarity and distributed again. This
is achieved by internal domains that are magnetized along the internal, transverse
easy axes. Those transverse domains can extend all through the volume, or they
can be connected to a basic domain wall so that the neighboring basic domain
is used to lead flux downwards. Because this system of compensating domains is
superimposed on the basic domains that would be present without misorientation,
these domains are called supplementary domains.
If a (moderate) magnetic field is applied along the surface-parallel easy axis,
Zeeman energy is added and those basic domains with magnetization along the field
direction will grow on expense of the opposite basic domains by 180◦ wall motion.
The rise in stray field energy, caused by the absence of oppositely magnetized
basic domains, is then compensated by an increasing number of supplementary
domains. At the same time those internal transverse domains, which are connected
to the basic domain walls, have to extend across the whole sheet thickness. So the
transverse domain volume is larger compared to the demagnetized state. This change
in relative domain volumes has consequences for the stress state of the sheet: as the
magnetostriction constant is positive for FeSi, the cubic crystal lattice is tetragonally
distorted along the magnetization direction. The basic domains thus cause an
408 R. Schäfer

N N
N S
N N
N N N N
Easy N N
axes S S N S
S N
S S S S

a) b) c)

S S S
S N N
N S S S
N N
N
N
N S S N
N N Tensile stress
S e)
S
d)

µ*-corrected
N N N
N
N
N
N

Without tensile stress With tensile stress


g) f)

Fig. 13 Interplay of magnetic energies, illustrated on the example of domain formation in


FeSi transformer sheets with (110)-related surfaces. Shown is the introduction of basic and
supplementary domains (b–d) in case of a slightly misoriented surface, starting from an ideally
oriented surface in (a). Tensile stress leads to domain refinement (e, f). In (g) the μ*-effect is
illustrated. (Image (d) is adapted by permission from Ref. [1] (c) Springer 1998)

elongation of the sheet along the rolling direction, while in the transverse domains
the sheet is transversely expanded. A change in the transverse domain volume will
thus result in a magnetostrictive change in length during remagnetization along the
[001] easy axis. Driven in a magnetic field at power frequency, the sheet will be
set in mechanical vibration leading to acoustic transformer noise. Furthermore, the
repeated destroying and rebuilding of supplementary domains forms an important
part of hysteresis loss as the energy bound in the supplementary domains is lost in
every cycle.
Magnetostrictive interaction can, however, also be favorably used in transformer
sheets. The supplementary domains are suppressed under tensile stress applied
along the preferred axis, because tensile stress magnetostrictively disfavors the
transverse domains that are attached to the supplementary domains. A domain state
as in Fig. 13c would thus result. Rather than superimposing supplementary domains
8 Magnetic Domains 409

to lower the stray field energy, which is forbidden now, a similar effect is achieved
by lowering the basic domain width (e). The domain images in (f) demonstrate this
effect. Obviously even ideally oriented grains assume a small domain width if they
are coupled to less well-oriented grains to achieve flux continuity. A narrow domain
with is favorable if the domains are excited by AC magnetic fields. The larger the
density of the walls, the smaller the velocity of every wall for a given induction level
which lowers domain wall-related eddy current effects (so-called anomalous eddy
current losses). In practice the tensile stress is created by the insulation coating that
is at the same time stress-effective. The planar stress exerted by the coating is for the
Goss texture equivalent to a uniaxial stress and will thus suppress the supplementary
domains.
Two further, energy-related aspects are worth to be noted: (i) so far it was
assumed that the domains are strictly magnetized along the easy crystal axes in the
demagnetized state and up to moderate applied magnetic fields. This is in fact true
for most of the volume domains. By approaching the (110) surface, however, the
magnetization bends toward the surface (Fig. 13g). So the surface poles are spread
over a certain volume and not just at the surface which helps to reduce the stray
field energy at the expense of some anisotropy energy, though. The phenomenon
is known as μ*-effect. (ii) The basic ±180◦ walls are zigzag folded across the
thickness as indicated in Fig. 13. Although the total wall area is larger than in case
of straight, perpendicular (110) walls that would have the smallest area, the total
wall energy is reduced by the folding. The reason is the specific wall energy, which
is lower for {100} wall orientations. The (110) wall therefore tends to rotate toward
these orientations, forming tilted or zigzag walls with a lower overall energy.

Domain Classification

The magnetic energy coefficients, listed in Fig. 10, can be combined in several ways
to obtain dimensionless parameters that reflect the interplay of energies and thus the
domain character. The ratio between anisotropy and stray field energy is the most
important. This ratio is called the quality factor, defined by

Keff
Q= . (8)
Kd

Here Keff is the effective anisotropy constant and Kd the stray field energy
coefficient defined in Eq. (5). If the anisotropy energy dominates over the stray field
energy (Q > 1), domains are formed that avoid an expense of anisotropy energy
while keeping the stray field energy as low as possible. If the stray field energy is
dominant (Q  1), stray fields are avoided by flux-closed domain patterns that
adapt to keep the anisotropy energy as low as possible. In the following discussion
we use the quality factor as primary criterion as it leads to the most fundamental
way of classifying domains and magnetic materials. In Fig. 14 a number of typical
materials are listed in the order of decreasing quality factor. Further criteria are the
410 R. Schäfer

0 0
µ0 Ms A K1 , Ku w w
Material Q 10
in Tesla in 10 J/m in J/m3 in 10 m in mJ/m2

1.7 107
SmCo5 1.05 12 39 0.84 57
hexagonal

4.9 106
CoPt (L10) 1.0 10 12 1.5 28
tetragonal

4.2 106
Sm2Co17 1.29 14 6.3 1.83 31
rhombohed.

4.5 106
Nd2Fe14B 1.61 7 4.4 1.25 23
tetragonal

3.2 105
BaFe12O19 0.48 7 3.5 4.68 6
hexagonal

4.5 105 11 = –45


Cobalt (Co) 1.79 31 0.35 8.3 15
hexagonal 44 = –260

4.8 104 100 = +22


Iron (Fe) 2.15 21 0.03 20.9 4
bcc 111 = –21

–4.5 103 100 = –55


Nickel (Ni) 0.60 8 0.03 42 0.8
fcc 111 = –23

Permalloy film 50 - 200


1.00 13 3 10 1 300
(Ni81Fe19wt%) Ku induced s

Fe74Cu1Nb3Si15B7 ~20
1.24 6 4 10 0.2 550 0.04
nanocryst. ribbon Ku induced s

Amorph. ribbon, ~3
0.6 2.5 2 10 0.1 900 0.01
Co-based Ku induced s

Fig. 14 Material parameters that are important for domain analysis. The listed materials are
ordered in terms of decreasing quality factor Q. Listed are furthermore saturation polarization
μ0 Ms , exchange stiffness constant A, first order anisotropy constant K1,u , magnetostriction
constant λ, wall width parameter Δ0w [see Eq. (14)], and specific wall energy of a 180◦ Bloch
wall γw0 [see Eq. (13)]. (Data are taken from Refs. [20, 21])

manifold of easy directions and the surface orientation of the investigated specimen,
which we treat as secondary criteria to classify the wide variability of domain
phenomena.
In Fig. 15 the interplay of stray field and anisotropy energy is illustrated by
comparing three material classes with uniaxial anisotropy but highly different
Q-factors. In all cases the easy axis is perpendicular to the plate surface, on
which domain observation was performed by Kerr microscopy, i.e., the specimens
are extremely misoriented with respect to the imaged surface. Compared are the
domains of a NdFeB single crystal (left column) with those in amorphous films and
ribbons (right column). The strong magnetocrystalline anisotropy of 4.5 · 106 J/m3
8 Magnetic Domains 411

D Ku

Dominating Dominating
anisotropy energy (Q >1) stray-field energy (Q<<1)
Film: Surface
D = 5 µm 5 µm
W
1
D =
S N S N 1 µm

Q=
7 µm N S N S 0.01
a) e)
Towards bulk:
N S N
7 N S N D =
25 µm

Q=
14 µm 0.0003
b) f)

D =
25 µm
40 µm
Q=
0.001
c) g)

D =
25 µm
120 µm
Q=
20 µm 0.002

d) h)
(Q 1)

polar magnetization in-plane magnetization

D =
1 mm

5 µm

Fig. 15 Classification of magnetic domains for three extrem cases of the quality factor. (a–d)
NdFeB single crystal, Q = 4.4. (e–h) FeBSi-based amorphous film and ribbons with stress-
induced perpendicular anisotropy and Q-values as indicated. (i) Cobalt single crystal, Q = 0.35.
In each case the domain images are taken on top surface, while the sketches show side views. The
Kerr images in (a–h) are adapted by permission from Ref. [1] (c) Springer 1998, while (i) was
obtained together with I. Soldatov, Dresden, and reproduced from Ref. [22] with the permission of
AIP Publishing
412 R. Schäfer

makes NdFeB a material with Q = 4.4, while Q  1 for amorphous material


owed to some weak (stress-)induced anisotropy around the order of 10–100 J/m3
(compare Fig. 14). An intermediate anisotropy of 4.5 · 105 J/m3 results in a quality
factor of 0.35 for hexagonal cobalt, shown on the bottom of the figure.
In case of the high-anisotropy material, the domains are strictly magnetized along
the easy axis to avoid anisotropy energy, even though this causes magnetic poles at
the surface thus costing stray field energy. This rule is strictly followed for films and
bulk specimens, though the domain character changes as function of thickness. For
films (Fig. 15a) a simple plate domain state with up-and-down domains is observed.
For rising
√ sample thickness D, the domain width W increases according to
W ∼ D. Beyond a critical thickness (about 5 μm in the example), the domain
walls get corrugated close to the surface (b), which leads to a better intermixing
of poles. This lowers the stray field energy compared to the hypothetical case of
straight domain walls. For further rising thickness (c), domain branching sets in:
close to the surface a fine domain pattern is enforced to minimize the stray field
energy by bringing opposite poles close together, whereas in the bulk, wide domains
are favored to save wall energy. With increasing thickness (d) further iterated
generations of domains are added, leading to a progressive domain refinement
toward the surface in a fractal way. Theory [1] yields a characteristic D 2/3
dependence of the basic domain width and a constant surface domain width within
the branching regime. As only up-and-down domains are involved in high Q
uniaxial material, the branching scheme is called two-phase branching.
Different arguments apply to the low-anisotropy material presented in the right
column of Fig. 15. In the limit of small thickness (not shown), a thin film with a weak
perpendicular anisotropy would be in-plane magnetized, because the anisotropy
energy density of this state would be smaller than the stray field energy density of a
uniformly perpendicularly magnetized state. Beyond a critical thickness, however,
the magnetization starts to oscillate out of the plane in a periodic manner to save
part of the anisotropy energy (Fig. 15e). The oscillation modulation assumes the
character of a two-dimensional flux-closed pattern, called dense stripe domains,
the half period of which typically equals the film thickness. Due to flux closure,
stray field energy is completely avoided as required for a Q  1 material,
whereas anisotropy energy and exchange energy are spent due to the deviations
from the easy axis and the non-parallel alignment of magnetization, respectively. For
larger thickness, an oscillating magnetization would consume increasing anisotropy
energy so that nature prefers a Landau pattern (f). Here the bulk is strictly
magnetized along the easy axis, and the expenses of exchange and anisotropy energy
are concentrated in regular domain walls and closure domains (being magnetized
perpendicular to the easy axis), respectively. When the perpendicular anisotropy
increases slightly, still within the Q  1 regime, the anisotropy energy in the
closure domains would rise so that beyond a critical anisotropy level (or beyond a
critical thickness
√ – note that the closure domain volume increases with thickness
due to a D-increase of the basic domain width) a three-dimensional branching
scheme takes over (g): the closure domains become themselves modulated in a
similar continuous manner as seen for films in (e), thus lowering their anisotropy
energy density. These stripe oscillations are connected with the basic domains by
8 Magnetic Domains 413

assuming the corrugated shape visible in the photograph and model. With a further
increase in anisotropy or sample thickness, the stripe domains do grow into regular
domains, the closure domains of which now decay into a further generation of
stripe pattern as shown in (h). This type of branching may be considered as multi-
phase branching which occurs despite the fact of having a uniaxial material rather
than cubic or other multiaxial materials. In any case, the overall domain patterns
(e - h) are completely free from stray field energy as required for such low-Q
material.
If anisotropy and stray field energy are competing with about equal magnitude
(Q ≈ 1), a hybrid structure is formed as shown in Fig. 15i for a thick cobalt
crystal. The character of the domain pattern is similar to that of NdFeB in the
macroscopic aspects, i.e., the branching mode in the bulk agrees with that of the
high-anisotropy material. The surface domain width, however, is smaller because of
the smaller wall energy of cobalt and theory predicts closure domains with a tilted
magnetization, owing to a lower anisotropy compared to NdFeB. By comparing
the two domain photographs in Fig. 15i, which show the out-of-plane and in-plane
magnetization components separately, it seems that the fine surface pattern of the
in-plane component resembles the branched domains of the amorphous ribbon in
image (h). Obviously the closure domains are modulated in a dense stripe domain
pattern. According to these findings, cobalt with its intermediate anisotropy forms
a kind of hybrid, following the high-anisotropy two-phase branching scheme in the
bulk and a low-anisotropy multiaxial branching scheme at the surface.
The classification principle discussed so far is generally valid – different crystal
symmetries or sample orientations just add modifications. Consider the case of
Fe3wt%Si material in which the three 100 axes are easy. The quality factor of 0.03
implies flux-closure domain configurations at zero field. Their character depends
on the surface orientation, and by having two further easy axis compared to the
uniaxial materials in Fig. 15, more degrees of freedom for the domain formation
are available. This becomes immediately apparent by looking at the domains of
the non-oriented sheet presented in Fig. 16a. The variety of flux-closed domain
patterns can be sub-classified according to the surface orientation: on an ideally
oriented surface, the principle of flux closure is immediately seen by wide domains
that are separated by well-oriented domain walls. Two grains with (110) and (100)
orientations are marked in the figure. A pole-free wall orientation requires that the
component of the magnetization perpendicular to the wall is the same on both sides
of the wall as indicated in Fig. 16b. For slight misorientation of a few degrees
flux collection is achieved by supplementary domains as seen for the (100)-related
surface in Fig. 16c (see Fig. 13 for a thorough explanation of this phenomenon for
a (110)-related surface). For stronger misorientation a domain branching scheme is
energetically preferred. In Fig. 16d this is illustrated for a (111) surface, i.e., the case
of extreme misorientation. Here in most of the volume a domain structure is formed
that occupies easy directions only, and these domains are joined so that no magnetic
stray fields are generated as required by Q  1. Near the surface zones, however,
the two requirements of using only easy directions and avoiding stray fields are
incompatible as the surface does not contain an easy direction. By branching nature
finds a compromise: the domains get finer toward the surface by adding several
414 R. Schäfer

~(111)

~(110)

(100) 10 µm e.a.

face
~(100) sur
1)-
(110) (11
100 µm
(100)-cut
a) d)

mn Pole-free
wall orientation
mn m2 requires:
m1 (m1 m2) n = 0
a)
n
b) c)

Fig. 16 The domains on a non-oriented Fe3wt%Si sheet (0.5 mm thick) reveal the influence
of the surface orientation on the domain character. (a) Low-resolution Kerr image giving an
overview. (b) Schematics of a 90◦ wall illustrating the condition for a pole-free wall orientation. (c)
Sketch of the fir tree structure, a supplementary pattern that appears on slightly misoriented (100)
surfaces of iron-like material. (d) High-resolution Kerr image on a (111) surface, together with
a schematics showing the phenomenon of multiaxial branching in iron-like material with cubic
crystal anisotropy. (The images in (a) are adapted by permission from Ref. [23] (c) Springer 2009)

generations of echelon domains in a fractal way (the number of generations depends


on the thickness). By the wide volume domains, wall energy is saved, and by the fine
surface domains, the volume of the outermost closure domains, which cannot be
magnetized along an easy axis, is minimized. Those closure domains are actually
magnetized in a continuously varying way similar to the dense stripe domains in
films (see Fig. 15e). So the unavoidable anisotropy energy of the surface zone is
reduced and right at the surface the magnetization lies parallel to the surface as
required by the low-quality factor.

Domain Walls

Domain walls form a continuous transition between neighboring domains. The


domain wall structure and character primarily depends on the Q-factor and thick-
ness of the specimen. Furthermore the wall angle (i.e., the relative angle of the
8 Magnetic Domains 415

Left domain (x)


1
sin (x)
Bloch path

x x 0
–4 –2 0 2 4
x / A/Ku

Néel S N
path
Right domain a) b) W180

Fig. 17 (a) Bloch and Néel wall paths in an infinite uniaxial material. The magnetic poles for the
Néel wall are indicated in the vector plot. (b) Wall profiles of a 180◦ Bloch wall. The indicated
wall width, W180 , is defined on basis of the slope of the magnetization angle ϕ(x). (Adapted by
permission from Ref. [1] (c) Springer 1998)

neighboring domain magnetizations) and magnetostriction may have an influence.


Here some basic aspects of domain walls are reviewed; for details and a thorough
review, we refer to Ref. [1].

Domain Wall Types

Two principle modes of magnetization rotation across a domain wall can be


distinguished: the Bloch and Néel wall. In Fig. 17a the two paths are illustrated
for the simplest of all domain walls, a planar 180◦ wall in a (hypothetic) infinitely
extended medium with uniaxial anisotropy that separates two domains of opposite
magnetization. If the domain magnetizations are parallel to the wall, there will be no
global magnetic poles, meaning that the component of magnetization perpendicular
to the wall is the same on both sides of the wall (compare Fig. 16b). In the Bloch
wall the magnetization rotates parallel to the wall plane, so there are no poles inside
the wall either (divm = dmx /dx = 0), and the stray field energy is zero. For the
Néel path the stray field energy would be maximum, making this wall path inferior
to the Bloch path in bulk material. Néel-type walls can nevertheless be favorable in
magnetic thin films as shown below.

Classical Bloch wall If magnetostriction and higher-order anisotropy constants are


neglected, the specific wall energy γw0 of a Bloch wall is written as an integral over
exchange energy (considering the non-parallel moments in the wall) and anisotropy
energy (due to deviations from the easy axis):
 ∞ π π
γw0 = [A(dϕ/dx)2 + Ku cos2 ϕ] dx , ϕ(−∞) = , ϕ(∞) = − . (9)
−∞ 2 2

Here x is the coordinate perpendicular to the wall, the angle ϕ rotates in the
wall from 90◦ to −90◦ , and ϕ(±∞) are the boundary conditions given by the
416 R. Schäfer

neighboring domains (see Fig. 17a). The solution of this ansatz is obtained by
variational calculus, which leads to a function ϕ(x) that minimizes γw0 under the
boundary conditions. Starting with Euler’s equation

2A(dϕ/dx)

= −2Ku sin ϕ cos ϕ , (10)

multiplying it with (dϕ/dx)


and integrating with respect to x leads to the first
integral:

A(dϕ/dx)
2 = Ku cos2 ϕ . (11)

According to this equation, the exchange and anisotropy energy densities are equal
at every point in the wall: at positions where the anisotropy energy is high, the
magnetization rotates rapidly leading to a large exchange energy. From Eq. (11) we
obtain

dx = A/Ku dϕ/ cos ϕ , (12)

which, inserted together with (11) into the total wall energy (9), yields
 ∞   π/2 
γw0 = 2 Ku cos2 ϕ dx = 2 AKu cos ϕ dϕ = 4 AKu . (13)
−∞ −π/2


Integration of (12) leads to the functional dependence sin ϕ = tanh(x/ A/Ku ),
which is plotted in Fig. 17b. From the indicated definition, the classical Bloch wall
width is derived as

W180 = π Δ0w , with Δ0w = A/K (14)

being the wall width parameter (see Fig. 14 for examples).


Although calculated for infinite samples, the classical Bloch wall also occurs in
“real” specimens if the quality factor of the material is larger than one. Then the
magnetic surface poles, which are caused by the out-of-plane magnetic moments
of the wall in case of in-plane magnetized domains, can be tolerated. For low
anisotropy material (Q  1), however, the requirement of pole avoidance enforces
different wall types. In Fig. 18 they are collected for materials with a (weak)
uniaxial, in-plane anisotropy Ku . There is a difference between thin films, thick
films, and bulk specimens.

Walls in thin films with Q < 1 From the micromagnetic point of view, magnetic
films are defined as “thin” if their thickness is below the classical Bloch wall width.
Then wall modes using a predominantly in-plane rotation of magnetization, known
as symmetric Néel wall and cross-tie wall, have a lower energy than the Bloch mode,
although magnetic poles cannot be avoided in those walls. The characteristics of
8 Magnetic Domains 417

D = 60 nm D = 460 nm
Asymmetric
Bloch wall
Cross-tie wall D = 10 nm
Symmetric
20 µm wall

Permalloy thickness in nm
50 100 200 375

Wall angle
Symmetric

S N Asymmetric
S S S N N N
S N
Tail Core Tail
Cross-tie Asymmetric
Symmetric Néel wall wall Bloch wall Vortex wall
//
1 2 3 4 5 6 7 10 12 15 20 25 30 40 50 60 5
D / A/K d D / A/K u
Side view

Cross Bloch line Circular Bloch line Swirl

Fig. 18 Phase diagram for various types of domain walls that exist in low-Q thin and thick films
at zero applied field and in a hard-axis field that causes magnetization rotation in the domains,
thus reducing the wall angle. The corresponding thicknesses for permalloy are indicated. Shown
are high-resolution Kerr images of permalloy together with sketches (symmetric Néel and cross-tie
wall, Bloch lines and swirl) and micromagnetically simulated vector plots (asymmetric Bloch and
Néel wall, calculated for permalloy). The contour lines in the calculated wall profiles indicate the
center of the walls, i.e., the surfaces on which the magnetization is strictly aligned in the drawing
plane. The pictures are taken and adapted from Ref. [1]. Since the anisotropy has only a moderate
influence on the wall energy in films, the diagram is valid for a wider range of low-Q materials.
(Adapted by permission from Ref. [1] (c) Springer 1998)

the symmetric Néel wall is its decomposition in a sharply localized core and two
extremely wide tails that take over a large part of the total rotation. A dipolar pole
pattern appears at the core, which carries about half of the pole density, and the other
half of the poles are displaced in the tails, both of them adding stray field energy.
The two characteristic lengths of a 180◦ Néel wall in a film of thickness D are then
given by

Wcore = 2 A/(Ku + Kd ) and Wtail ≈ 0.56DKd /Ku . (15)

Because Kd is much√larger than Ku for materials with Q  1, the core width


roughly scales with A/Kd , i.e., in the core the exchange energy (A) is primarily
balanced by the stray field energy (Kd ) that is connected with the magnetic poles.
The tail is rather determined by a balance between stray field (Kd ) and anisotropy
418 R. Schäfer

energy (Ku ). The stray field energy thus has an opposite effect on both parts of
the wall. The poles in the extended tails of symmetric Néel walls lead to strong
wall interactions as soon as the tails overlap. In thin permalloy films, for example,
this happens up to wall distances as high as 100 μm. This interaction can result
in attraction or repulsing of neighboring walls if the walls have opposite or equal
rotation senses, respectively, called unwinding and winding walls.
The pole density of a symmetric Néel wall can be drastically lowered by reducing
the wall angle, e.g., by applying a magnetic field perpendicular to the easy axis. If
the total√integrated pole density in each half of the wall is μ0 Ms , it is reduced to
(1 − 1/ 2)μ0 Ms D for a 90◦ wall. As the stray field energy varies quadratically
with the pole density, a 90◦ Néel wall has just 12% of the specific energy of a 180◦
wall, leading to strong preference for lower-angle Néel walls. As a consequence,
the symmetric 180◦ Néel wall at zero field is replaced in thin films by the cross-tie
wall, a composite wall that actually consist of crossing 90◦ walls. Although the total
length of the wall segments within this pattern is 3 to 4 times larger than the original
180◦ wall area, the total wall energy is smaller. Below a critical wall angle of about
135◦ the cross-tie wall is replaced by the symmetric Néel wall.
For very thin films (below about 50 nm for permalloy), the cross-tie wall becomes
unstable, however. The reason are the Bloch lines occurring for topological reasons
at the centers of the continuous transitions between the partial walls. Depending
on the magnetization environment, circular and cross Bloch lines (also known as
vortex and antivortex) are distinguished. Bloch lines are not singularities, because
the magnetization turns perpendicular to the surface when approaching their center.
As mainly stray field and exchange energy are √ competing in a Bloch line, its core
width scales with the characteristic length A/Kd (about 5 nm for permalloy),
which is small compared to the wall width. Bloch lines are therefore strongly pinned
at defects of similar size, thus acting as a source of wall coercivity in thin films. With
decreasing film thickness, the stray field energy of a Bloch line increases, so that for
very thin films the symmetric Néel wall is energetically preferred again despite its
higher specific wall energy. Bloch lines are present in such walls but much less
frequently compared to cross-tie walls.

Walls in thick films with Q < 1 Films with√a thickness beyond the Néel-Bloch
wall transition, which occurs typically at 20 A/Kd or ∼100 nm for permalloy
films, are defined as thick films. Asymmetric Bloch walls, favored for small or zero
field, are characteristic for such films. Due to a two-dimensional internal vortex
structure, which appears asymmetric relative to the wall plane, this wall mode is
completely stray field-free although it appears like a common Néel wall at the
surface. The eddy is geometrically √ determined so that the wall width scales with
the film thickness rather than with A/Kd as for the symmetric Néel wall. Because
the anisotropy energy plays a minor role (Q < 1) and stray field energy is excluded,
the exchange energy is the predominating term for the asymmetric Bloch wall. The
wall has four equivalent orientations, since the vortex can be on the left or right and
the center magnetization can be upward or downward. Usually all four orientations
appear within the same wall with the segments being separated by Bloch lines.
8 Magnetic Domains 419

The structure of the asymmetric Bloch wall loses its mirror symmetry in an
applied hard-axis field (i.e., for reduced wall angles). Beyond a certain field value,
another vortex wall mode, the asymmetric Néel wall, becomes favorable because it
is better adapted to the applied field. It is point-symmetric in the cross section, the
horizontal magnetization component is given by the field direction thus pointing in
the same direction at both surfaces, and the out-of-plane excursion in the volume
can be either up or down leading to two equivalent variants. Like the symmetric
Néel wall, the asymmetric Néel wall also splits off an extended, logarithmic tail but
with much less pole density as most of the poles are avoided by the two dimensional
vortex pattern [25]. Below a critical wall angle of ∼ 45◦ the asymmetry disappears
in favor of the symmetric Néel wall.

Walls in√bulk specimens with Q < 1 The upper limit for thick films can be set at
about 5 A/Ku for uniaxial films (or 1.5 μm for permalloy), when the characteristic
thick-film Bloch wall gradually transforms into the classical Bloch-type wall at
least inside the sample, whereas the vortices are confined to the neighborhood of
the surfaces, thus providing flux closure and pole avoidance. At the surfaces the
magnetization of such vortex walls rotates parallel to the surface (reminiscent of the
Néel mode and sometimes called a Néel cap), and the surface wall width is wider
than the width of the Bloch part in the volume that is given by Eq. (14). If two wall
segments with opposite surface rotation sense meet, a Bloch line is enforced for
topological reasons. In its core the magnetization
√ turns perpendicular to the surface
forming a swirl with a diameter of the order of A/Kd rather than a singularity.

Walls in specimens with Q > 1 For Q  1 materials like permalloy or


amorphous ribbons, the vortex wall in (in-plane magnetized) bulk specimens is
completely pole-free, while it becomes less so when Q increases. The Q limit for the
occurrence of a surface cap is around one, and beyond that limit, the classical Bloch
wall with surface poles is preferred. Bloch walls are also favored in high-Q materials
with perpendicular anisotropy where they separate band or bubble domains. The
statics and dynamics of such walls together with their Bloch line substructures were
thoroughly investigated back in the 1970s in connection with the development of
the magnetic bubble memory [24]. Micrometer-thick magnetic garnet films with a
quality factor of 3 were used for that purpose.
In more recent years also the existence of Néel walls is discussed for high-Q
materials. In ultrathin mono- or multilayer films with perpendicular anisotropy, spin-
orbit coupling effects may lead to a surface- or interface-induced Dzyaloshinskii-
Moriya interaction (DMI, compare section “Driving Forces for Domain Forma-
tion”). This effect is pronounced at the interface between ferromagnetic and heavy
nonmagnetic metals, with Pt/Co bilayers being a typical example. If in a regular
Bloch wall the magnetization may rotate in a left- or right-handed way as a result
of the symmetric exchange interaction, the DMI can lift the left/right degeneracy
thus leading to a homochiral Néel wall structure [26] as indicated in Fig. 19a. The
Néel walls have a specific chirality that is fixed by the sign of the Dzyaloshinskii
constant, which corresponds to the exchange constant A for Heisenberg exchange.
420 R. Schäfer

Ferromagnetic
layer
a) Spin-orbit layer

Half hedgehog vortex Half antivortex

Swirl,

vortex

b) c) Half antivortex d) Half antivortex

Fig. 19 (a) Néel walls of the same chirality, induced by Dzyaloshinskii-Moriya interaction.
Shown is the cross section through a bilayer film. Head-on domains in an extended film (b) and
in Q  1 nanostrips of rectangular cross section with transverse (c) and vortex wall (d), seen
from the top in each case. The latter are of lower energy in wider strips and thicker films. (The
schematics in (c) and (d) are redrawn following Ref. [27])

Special walls The full development of cross-tie walls or extended Néel tails
requires sufficiently extended films. If geometrical restrictions are present, specific
wall structures may develop. A prominent example are head-to-head or tail-to-tail
walls in low-anisotropy (e.g., permalloy) film strips with sub-micrometer width,
which have been of interest in the early stages of research on current-induced
wall displacement devices (section “Current-Driven Domain Wall Motion”). If two
domains meet head-on in an extended film, a straight transversal wall would have
strongest pole concentration. By forming a zigzag shape (Fig. 19a), the pole density
decreases at the expense of wall surface, though. Depending on the film thickness,
the core of this wall may be of Néel, cross-tie, or Bloch-type, and part of the
poles are distributed in tails on both sides similar to the tails of the symmetric
Néel wall. As the zigzag amplitude and period require spatial ranges between tens
and hundreds of micrometer, zigzag walls are not possible in sub-micrometer wide
strips. Special wall types, termed transverse and vortex walls, are rather formed
(Fig. 19b, c). Their width is determined by the shape anisotropy of the wire and
thus varies approximately with the wire width. Both structures are characterized by
topological objects as marked in the figure. For details see the reviews [28, 27].

Domain Wall Dynamics

If a domain pattern is exposed to a magnetic field that favors certain domain phases
against others, a pressure acts on the domain walls which separate these phases,
and they are set in motion in order to lower the energy of one domain phase with
respect to another. The wall mobility and velocity are determined by a number of
mechanisms:
8 Magnetic Domains 421

• Domain walls may interact with inhomogeneities in the material like nonmag-
netic inclusions, grain- and phase boundaries, stress centers, etc. if the wall saves
energy when resting at these positions. For example, a nonmagnetic inclusion
exerts a pinning force to the wall as the wall saves area and thus energy when
it intersects the inclusion. The pining force is described by the balance between
the gain in Zeeman energy when the wall moves a certain distance, driven by the
field, and increase in wall energy. When some critical field, the static coercivity
Hcs , is exceeded, the wall breaks away from its pinning site and starts moving,
typically by a Barkhausen jump till it is pinned by the next obstacle. This leads
to micro-hysteresis steps that sum up to the macro-hysteresis curve. If thermally
activated creeping effects are relevant (like for the films in Fig. 3), Hcs will be
lowered with rising temperature.
• Once the wall is set in (steady) motion, it moves with a velocity v that is
proportional to the driving field Hext according to v = μw (Hext − Hcs ) where μw
is the wall mobility. It is determined by eddy currents, intrinsic damping, or other
dissipative processes, which all act like friction to the moving wall and additively
determine the overall mobility.
• One of the origins of wall friction may be thermal diffusion leading to an
induced anisotropy. If the wall moves fast enough so that the induced anisotropy
changes little when the wall moves by its width, the diffusion effects just add
to friction-like losses. If the wall rests, however, it becomes hard to move after
some time, resulting in static coercivity. In between, a fractional decrease of
wall mobility (i.e., permeability) with time under otherwise constant operation
conditions, known as disaccommodation, is the consequence. The classical
example for such a process is the diffusion of carbon in iron. In metallic materials,
disaccommodation can be avoided by removing or binding the disturbing atoms
or point defects. In high-permeability ferrites, on the other hand, electron hopping
between two- and three-valent iron ions on the same octahedral lattice sites may
lead to disaccommodation effects that are hard to suppress.
• While disaccommodation may be important for slow magnetization processes,
eddy current effects in metallic materials become more effective with increasing
frequency. For a moving domain wall, the flux change in the region swept out by
the wall induces a voltage according to Faraday’s induction law, which causes
the flow of eddy currents. They are responsible for a magnetic eddy current
field that opposes the applied field. Because the effective field acting on the
wall is consequently less than the applied field, the wall velocity is less than
if the eddy currents did not exist, i.e., the wall motion is damped. For a planar
wall oriented perpendicular to the surface of a sheet, for example, the eddy-
eddy
current-limited mobility is μw = v/H = 3 · 106 /(Ms σ D). The lower the
electrical conductivity σ and the smaller the thickness D, the more restricted are
the eddy currents and the larger the mobility. Due to eddy currents a moving
wall is retarded more at the center than at the surfaces, which may lead to drastic
wall deformations. At high velocities the walls become grossly deformed to run
mostly parallel to the surface. Eventually, magnetization processes occur mostly
in a surface-parallel sheet, a phenomenon known as the skin effect.
422 R. Schäfer

• If disaccommodation and eddy current effects are absent or negligible, which is


the case in clean non-conducting media and in metallic thin films, the domain
wall dynamics is dominated by the gyrotropic nature of the magnetization vector
that is generally described by the Landau-Lifshitz-Gilbert-Slonczewski (LLGS)
equation

dm dm
= −γ m × H eff − α m × +τ, (16)
dt dt

where γ = μ0 gμB /h̄ is the gyromagnetic ratio in units [m/As] with μB the
Bohr magneton and g the Landé factor. The LLGS equation describes the
gyromagnetic precession of the magnetization around an effective field Heff
(first term) and its relaxation toward Heff (second term) as illustrated in Fig. 20
(for the τ -term, denoting current-induced torques, see section “Current-Driven
Domain Wall Motion” and Fig. 24c). The effective field incorporates applied and
demagnetizing fields, but also contributions from anisotropy, exchange stiffness,
etc. The second term, also called viscous-damping term, is caused by intrinsic
dissipative phenomena like the scattering of spin waves on lattice defects or the
relaxation of magnetic impurities, which are considered in the dimensionless
damping factor α. At low frequencies the loss term in the LLGS equation is
predominant, and the magnetization turns toward the effective field until both
vectors are parallel within some finite time. The gyromagnetic term has to be
taken into account only in the GHz frequency regime where it leads to strong
precessional effects on moving domain walls.
As an example let us discuss a one-dimensional Bloch wall in an extended film
with uniaxial anisotropy perpendicular to the film plane, exposed to a magnetic
field along one of the domain directions, i.e., perpendicular to the magnetization
in the center of the wall (Fig. 21a, b). If the field is slowly increasing, the damping
term in the LLGS equation dominates and the applied field exerts a direct torque
on the spins inside the wall as illustrated in (a). When the applied field is
sufficient to overcome pinning forces, the torque causes the wall to move in x-
direction. If a step-function drive field is applied to a stationary wall, however,
the wall is set in fast motion, and gyrotropic precession will become dominant

Fig. 20 Illustration of the Heff


first two terms of the Landau-
Lifshitz-Gilbert-Slonczewski dm
equation. Indicated is the mx
dt
gyrotropic precession of the
magnetization vector around m x Heff
the effective field and its
relaxation toward Heff m
8 Magnetic Domains 423

Hext

Wall velocity
Walker +
(slow) break-
down
Time

Average wall velocity


e)
e
qu

on
r
To

oti
ll m
Chaotic
x tion
Wa l mo
a) motion
na
ce ssio
Pre

H ext Conservative,
S
(fast) Hd m steady motion
N
m H ext d)
b) m Hd
z z

x
y y

c) Bloch wall at rest Moving Bloch wall

Fig. 21 Torques on the magnetization vector of a 180◦ Bloch wall in case of a slowly (a) and
fast (b) changing magnetic field. The development of a Néel component in a moving Bloch wall is
illustrated in (c). The wall velocity as a function of magnetic field (d) is characterized by the three
typical regimes of steady, chaotic, and precessional motion. In the latter mode, the wall oscillates
between forward and backward motion as indicted in (e)

as visualized in (b). Rather than directly inducing a rotation of the magnetization


vectors inside the wall that would give rise to wall motion, the pulse field Hext
firstly generates a rotation into the direction perpendicular to the wall plane
according to the Zeeman torque m × H ext in (16), thus inducing a deviation from
the Bloch path (ϑ = 0, compare Fig. 17a). This canting of magnetization out of
the wall plane results in the appearance of free poles that generate a transversal
demagnetizing field H d , which then leads to the required in-plane precession
according to the demagnetizing torque m × H d , finally setting and keeping
the wall in motion. The transverse magnetization component causes stray field
energy that can be considered by replacing Ku in the static wall equations by
K = Ku + Kd sin2 ϑ for the moving wall, leading to
 
γw = 4 A(Ku + Kd sin2 ϑ) and Δw = A/(Ku + Kd sin2 ϑ) (17)

for the wall energy and wall width parameter. In case of dynamic equilibrium, the
Zeeman torque is cancelled by the demagnetizing torque and damping, so that the
wall translates with a constant ϑ and a fixed spin structure throughout the wall
424 R. Schäfer

as illustrated in Fig. 21c. During this “conservative” (i.e., constant-velocity) or


“steady-state” (i.e., the wall structure does not change) mode of wall motion, the
wall velocity v is found to increase linearly with the applied field amplitude Hext ,
corresponding to viscous motion according to the Walker relation:

v = γ Ms Δw sin ϑ cos ϑ = (γ /α) Δw Hext . (18)

As ϑ increases with rising velocity [see first term in (18)], the wall width
decreases, and the wall energy increases according to (17). From v = μw Hext
and (18), the wall mobility of a 180◦ wall is derived as μw = 4Aγ /(αγw ). The
higher the energy of a wall, the lower its mobility.
With increasing field the velocity increases up to a maximum value that is
reached when the equilibrium between damping/demagnetizing torque and drive
field torque is broken, i.e., the Zeeman torque begins to dominate. The peak
velocity vp and peak field Hp are found as [1]

 
vp = γ 2AQ/μ0 f (Q) and Hp = αHa f (Q) 4 1 + 1/Q (19)


with f (Q) = 1 + 1/Q − 1. The peak field, also called Walker field, is
obviously proportional to the anisotropy field Ha = 2Ku /(μ0 Ms ). Beyond the
peak velocity, a steady-state motion with a fixed spin structure is not possible
anymore and a transient region of “negative mobility” is found (Fig. 21d) in
which the average velocity decreases with increasing applied field. The wall
develops spatially inhomogeneous modes in that regime with parts of the wall
advancing, while other parts are lacking back in an inhomogeneous, “chaotic”
way. For the following linear regime, the wall angle ϑ undergoes a more or
less free precession with the wall magnetization cycling periodically through
Bloch and Néel wall configurations, while the wall itself oscillates back and
forth as indicated in Fig. 21e. The motion is oscillatory as the torque term is
periodic with respect to ϑ. Due to damping, the backward motion is smaller than
the forward one, resulting in a positive average displacement with a non-zero
average velocity. Anyway, domain wall motion beyond the peak velocity of the
external field has to be considered highly irregular. For this reason the maximum
velocity possible for a stationary wall motion is also called the “breakdown”
velocity, and the phenomenon is known as Walker breakdown. It determines the
maximum possible speed of devices that are based on domain wall motion. In
(thick) bubble garnet films, for instance, the peak velocity reached values of
the order of 100 m/s. Compared to extended and thick films, the Walker field
is reduced in case of confined systems like nanowires or ultrathin films due to
complex demagnetizing field effects [29].
The increase in wall energy up to the breakdown velocity can be interpreted as
if the wall exhibits an effective (not a real) mass. The wall mass becomes obvious
from the equation of motion of a vibrating wall
8 Magnetic Domains 425

d2 x dx
mw 2
+ b + ax = 2Ms (Hext − Hcd ) , (20)
dt dt

which is analogous to the equation of motion of a (real) mass on a spring and


in which x is the wall normal displacement coordinate. The right hand side of
(20) is the force per unit area or pressure that the alternating magnetic field
exerts on the wall. The second term on the left is called viscous damping,
representing a resistance proportional to the wall velocity that is caused by
the mentioned disaccommodation, eddy current, and intrinsic damping effects.
The damping coefficient b is indirectly proportional to the mobility μw . The
third term represents the mentioned force due to pinning effects with the
restoring coefficient a. The first term, i.e., the product of mass mw and wall
acceleration, finally represents the inertia of the wall, caused by the resistance
of the spins to sudden rotation due to gyrotropic precession. To derive the wall
mass, firstly consider the demagnetizing field due to the out-of-plane rotation
of the magnetization vector in the moving wall, which is given by Hd = −Mx
according to Eq. (4, with N = 1). This field acts on each vector in the wall and
induces the mentioned precession in the wall plane that finally leads to the wall
displacement along x as illustrated in Fig. 21b. The rotational velocity of this
Lamour precession is given by dϕ/dt = γ Hd = −γ Mx . At the same time,
it is related to the wall velocity v by dϕ/dt = (∂ϕ/∂x)(dx/dt) = (∂ϕ/∂x)v. A
comparison of these equations leads to Mx = (v/γ )(∂ϕ/∂x). The demagnetizing
field leads to stray field energy which√can be seen as an additional kinetic term
γwdem to the static wall energy γw0 = 4 AKu . According to (4), and by using (11)
and (12), the stray field term is given by

  
1 ∞ μ0 v 2 Ku π/2 μ0 γw0 2
γwdem = − μ0 Mx Hd dx = cos ϕdϕ = v . (21)
2 −∞ 2γ 2 A −π/2 4γ 2 A

This energy is proportional to v 2 , and by expressing it classically as 12 mw v 2 , the


Döring mass of a 180◦ wall,
1
mw = μ0 γw0 /(Aγ 2 ) = 2μ0 /(Δ0w γ 2 ) , (22)
2

is derived. The mass is inversely proportional to the wall width parameter Δ0w .
For a 180◦ wall in iron with γ = 2 · 105 m/As, A = 21 · 10−12 J/m, and γw0 =
4 · 10−3 J/m2 (see Fig. 14), for example, the wall mass is approx. 3 · 10−9 kg/m2 .

Thorough reviews on wall dynamics can be found in the textbooks by Malozemoff


and Slonczewski [24], Hubert [30], and the review articles by Filippov [31] and
Thiaville and Nakatani [28] with emphasis on vortex walls and walls in confined
nanostrips, respectively.
426 R. Schäfer

Current-Driven Domain Wall Motion

Conventionally, domain walls are set in motion by the pressure that acts on a
wall by an applied magnetic field according to Eq. (20). An alternative way is
wall motion driven by spin-polarized current. In small structures like magnetic
nanowires this mechanism may dominate over the effects due to the self-field of
the same current. While field-driven wall motion would be opposite for neighboring
domain walls (domains either expand or shrink in a magnetic field), all walls are
pushed in the same direction by polarized current pulses making the racetrack
memory concept [8] possible (see Fig. 7b). Current-driven domain wall motion
is based on the phenomenon of spin-transfer-torque (STT) that was predicted by
Berger and Slonczewski [32,33,34]: because of spin-dependent diffusive scattering,
an electrical current carries spin angular momentum when it travels through a
magnetic material with the majority of the electron spins being aligned along the
magnetization direction. As these spin-polarized itinerant electrons pass through a
domain wall, the electron spin precesses and changes its polarization direction due
to the exchange interaction with the localized spins of the wall. To conserve the total
spin, the change of momentum of the traversing electrons needs to be transferred to
the wall spins. The local magnetic moments in the wall thus experience a torque that
changes their direction, possibly resulting in a displacement of the wall along the
direction of the electron flow.
Phenomenologically, the SST effect is taken into account by the τ -term in the
LLGS equation of motion (16). For arbitrary current direction and wall structure,
this term is written as the sum of two spin-torque terms proportional to the gradient
of the magnetization [35]:

γ h̄Pj
τ = −(u · ∇)m + β m × [u · ∇)m] , with | u |= . (23)
2eMs μ0

The term (u · ∇)m reads u(∂m/∂x) for a current along the x direction. The vector
u scales with the current density j and represents the maximum wall velocity along
the direction of electron flow under the assumption that the conduction electron
spin moments are fully converted into wall displacement. The dimensionless
parameter P represents the spin polarization of the current with 0 < P < 1
(around 70 % for permalloy), and e is the electron charge. The larger P , the more
efficiently the STT can move a domain wall. The first and second terms in (23)
describe adiabatic and non-adiabatic STT contributions, respectively. The adiabatic
(or Berger) term is derived under the assumption of the mentioned conservation of
the spin angular momentum, i.e., the entire spin angular momentum of the injected
carriers is transferred to the magnetization. In this limit the spin of the incoming
conduction electrons follows the magnetization profile of the domain wall. This
assumption is valid for sufficiently wide walls, for which an electron as enough
“time” to adapt to the localized magnetization while precessing and exchanging
spin angular momentum with the localized moments. If the magnetization gradient
becomes large, however, the assumption of adiabatic propagation is no longer
8 Magnetic Domains 427

valid. Then the spin polarization of the current cannot follow the rapidly changing
magnetization profile anymore, i.e., the exchange of spin angular momentum
between the conduction electron and the local moment is not conservative, leading
eventually to mistracking or even reflection (domain wall scattering) of the itinerant
spin. These effects are considered in the second, non-adiabatic spin torque term
in (16) that was independently introduced in refs. [36, 35] and which describes a
torque perpendicular to the adiabatic torque. The deviation from adiabaticity of the
STT is designated by the non-adiabatic STT parameter β, the ratio between the
non-adiabatic and adiabatic torques. It considers the mentioned spin mistracking
and spin relaxation in the wall by spin-flip scattering at impurities, phonons, etc. In
general both adiabatic and non-adiabatic terms have to be considered for a given
domain wall structure.
For current-driven wall motion, similar mechanisms are predicted as for the field
torque. Their appearance depends on the relative values of the Gilbert damping
parameter α and the β parameter, which both are dimensionless. As schematically
shown in Fig. 22a, there is some threshold current density in the absence of the non-
adiabatic torque (β = 0), below which the wall does not move at all. This threshold
is intrinsic (i.e., not caused by extrinsic pinning effects) because it originates only
from magnetization changes in the wall. The relevant torques for this case are
illustrated in Fig. 22b for the example of a 180◦ Bloch wall, which is reduced
to its central vector like in Fig. 21b. The adiabatic torque τAD on this moment,
together with the resulting demagnetizing fields, leads to a Néel-like deformation
of the wall structure and to a downward torque on the central vector that effectively
compensates τAD . This compensation of torques prevents wall motion up to a critical
velocity (i.e., current density) uc = γ Δ0w Ha /2 that is proportional to the anisotropy
field Ha . Up to uc , the adiabatic torque thus acts like an applied magnetic field
transverse to the wall plane that just distorts the wall without inducing a steady
motion. Above the threshold, the internal torque is not sufficient anymore to balance
the spin transfer torque, and the wall is set in motion together with a continuous
precession of the wall magnetization.
Wall velocity

Steady Precessional
motion motion

> m AD H d,1 m
Hd
=0 H d,1 m
2

NA
d,

=
H

0 H d,2 m Hd m
a) uc u b) c)

Fig. 22 (a) Average wall velocity as a function of the effective drift velocity u of the spin current,
schematically plotted for three different non-adiabaticity coefficients β (Adapted from Ref. [35]).
The adiabatic torque leads to wall distortion and motion (b), while the non-adiabatic torque (c)
just moves the wall. Indicated are the primary STT vectors, τAD and τNA , together the following
torques due to the generated demagnetizing fields [37]
428 R. Schäfer

The non-adiabatic torque (Fig. 22c), on the other hand, causes a demagnetizing
field with a consequent downward torque that sets the wall in steady motion right
away by mimicking an easy axis field (compare Fig. 21b). According to Fig. 22a
this is true for any DC applied current if β = α, and it is true for currents smaller
than a critical current if β > α. In the steady-wall motion regime, the non-adiabatic
torque together with the damping constant α dictates the velocity v = βu/α. It is
obtained by replacing Hext in (18) by the spin transfer equivalent field βu/(γ Δw )
[29]. Beyond the critical current, the adiabatic term dominates, and the wall slows
down (v = u) due to the onset of precession and backward motion accompanied
by periodic transformations between Bloch and Néel wall. The critical current
thus corresponds to the Walker breakdown of the field-driven dynamics (compare
Fig. 21d). So the non-adiabatic spin transfer plays a crucial role in current-induced
wall motion: it allows for a non-zero mobility of the wall in a steady-state, viscous-
flow regime with the mobility proportional to β/α, resulting in the elimination of
the intrinsic threshold current even for small values of β. Any threshold current is
then only determined by extrinsic pinning.
Compared to relatively wide domain walls in in-plane magnetized nanowires
(see Fig. 19), which can readily expand to many times their equilibrium size under
current torques [38], the domain walls are of Bloch-type in nanowires with an
anisotropy perpendicular to the plane of the wire. Well studied are atomically
engineered cobalt-nickel superlattices in which interface anisotropies dominate over
the volume terms (anisotropy and magnetostatic), thus leading to the perpendicular
easy axis. With typical wall widths of some nanometers, the magnetization gradient
is large, and consequently a high non-adiabatic effect is to be expected. This makes
such materials attractive for current-induced domain wall motion as it pushes the
walls like the torque of a magnetic field with an increased STT efficiency compared
to in-plane domain walls. Wall velocities of the order of 100 m/s were measured in
such films, comparable to the speeds found for field-driven wall motion in bubble
material.
The spin transfer torques, described so far, are volume effects. Even higher
current-induced wall velocities (before Walker breakdown) in perpendicular-
anisotropy media can be achieved by making use of interface-related spin
torque effects that are obtained by depositing the magnetic stripes adjacent to a
nonmagnetic conductive layer with strong spin-orbit interaction (SO layer) like
heavy metal platinum, palladium, iridium, or tungsten films (see refs. [39, 38] for
reviews). The following current-induced SO-torques may contribute:

• Spin Hall effect (SHE): If a charge current passes the SO layer, the SHE creates a
spin current due to spin-orbit scattering of the conduction electrons. The spin
current flows in a direction perpendicular to the charge current toward each
surface of the SO layer. It may thus be seen as pure spin current, i.e., without
any charge flow in the perpendicular direction. The direction of the spin is
perpendicular to both the charge and spin currents as illustrated in Fig. 23a. The
interfacial spin accumulation is injected into the adjacent magnetic layer where
it exerts a STT [39]
8 Magnetic Domains 429

No SH Heff
Heff
Bloch
wall

SH
Heff
Heff
Néel
wall
tion
Wall mo
a) b)

Fig. 23 (a) Illustration of the Spin Hall Effect and its torque τSH . Indicated are the perpendicular
domain magnetizations and the central vectors for both rotation senses of a Bloch- (upper-) and
Néel wall (lower panel). (b) The Spin-Hall torque on a Néel wall leads to an effective magnetic
field that depends on the wall rotation sense, leading to a wall motion along or opposite to the
charge current direction. For each sketch, the spin-orbit layer is indicated as bottom layer, being in
contact with the ferromagnetic film on top

τSH ∼ m × σ × m . (24)

Here σ = j × z with j and z being the unit vectors of current and out-of-plane
direction, respectively. The spin-Hall torque τSH manipulates the magnetization
in the magnetic film as shown for domain walls in Fig. 23a. For Bloch walls of
either rotation sense, the torque vanishes, while for Néel walls a transverse torque
acts on the wall spins that points in the same direction (along the injected electron
spin direction) for both rotation senses of the wall magnetization. The resulting
demagnetizing fields finally lead to an effective magnetic field acting on a Néel
wall. As illustrated in Fig. 23b, this field is opposite for Néel walls with opposite
rotation senses. Consequently, neighboring walls will be pushed in the same
direction if they have the same chirality, i.e., opposite directions of their central
vectors. Such homochiral Néel walls can be enforced in ultrathin multilayers with
interface-induced Dzyaloshinskii-Moriya interaction; see section “Domain Wall
Types” and Fig. 19a.
• Rashba effect: This spin-orbit-related phenomenon, reviewed in Ref. [40], may
be induced by an electric field that is mediated by the symmetry breaking at
the interfaces of a typical multilayer composed of heavy-metal/ferromagnetic
metal/oxide films and that is oriented perpendicular to the film plane. When an in-
plane charge current flows in the magnetic layer, the itinerant electrons transform
the electrical field into a (Rashba) in-plane magnetic field that is experienced by
the flowing electrons and that polarizes them. By exchange interaction they can
then exert a field-like SST on the local magnetization of the magnetic film which
may result in some pressure on the domain walls [41, 39]. In contrast to the Spin
Hall effect, which is a bulk effect, the Rashba effect is an interface property.

The direction of domain wall motion and their velocity are determined by
the subtle interplay of these phenomena, which can be tuned independently by
430 R. Schäfer

varying the thickness of the SO layers and the composition of the ferromagnetic
layer. Current-induced domain wall motion with velocities of several 100 m/s
can be achieved by making use of SO effects. Still higher wall velocities of
almost 1.000 m/s were measured in nanowires that are formed from synthetic
antiferromagnets, i.e., from two antiferromagnetically coupled films via an ultrathin
nonmagnetic spacer layer [42, 38]. The high velocity is attributed to the vanishing
net magnetic moment in such film systems.
Besides the so far discussed nanowires with the current flowing in the plane,
STT [34] manifests itself also in sub-micrometer diameter pillars, composed of
magnetic multilayers and flooded by a current perpendicular to the plane. Such
pillars are the basic units of the SST-MRAM (magnetic random-access memory)
concept. Magnetically, they can be well described in the framework of a macrospin
model, meaning that the magnetic films of the pillar are assumed to move in a single-
domain state rather than by short-wavelength modes. As domains thus do not play a
role, we will touch this case just briefly here, leaving details to specialized literature
(see refs. [43, 44] for reviews).
Assume a ferromagnet/normal-metal/ferromagnet trilayer system as illustrated
in Fig. 24a, b. The two ferromagnetic layers of this spin valve-type device are
depicted as “fixed” and “free,” indicating that they are less and more susceptible,
respectively, to the STT. If unpolarized electrons are entering the fixed layer from
the left (Fig. 24a), they will undergo spin-filtering, i.e., by exchange interaction with
the magnetic moment of the fixed layer, the flowing electrons have an averaged spin
moment parallel to the magnetization of the fixed layer when emerging into the
nonmagnetic spacer film. If the spacer is thinner than the spin diffusion length, this
current will remain spin-polarized when it enters the free layer, in which the spin
components of the incoming electrons transverse to its magnetization are absorbed.
As the spin angular momentum is conserved, this lost transverse momentum of the
electrical current must be absorbed by the magnetic film, leading to a torque that

ted d Spin transfer Heff , m


smit itte torques for:
Tran ctrons r a nsmrons
ele T lect 0
Free e I<
layer 0
Fixed I>
layer
m ted
ns e flecrons
ctro R ct
m free
ele ele

a) b) c)

Fig. 24 (a, b) Illustration of the STT in a spin-valve-type trilayer, leading to the switching of the
free layer magnetization either parallel (a) or antiparallel (b) to that of the fixed layer (starting
from an assumed perpendicular initial configuration, with the thickness of the nonmagnetic spacer
layer exaggerated for visualization purpose). The magnetization of the fixed layer is stabilized by
making it thicker than the free layer, by using a material with a larger total moment or by exchange
coupling to an antiferromagnetic layer. (c) Schematics of the STT in analogy to Fig. 20
8 Magnetic Domains 431

tends to turn the free layer magnetization toward the orientation of the incident spin
polarization. So the parallel alignment of magnetization in the fixed and free layer
is stabilized. This STT requires that the moments of the two layers are initially
non-collinear, perhaps due to thermal fluctuations (exact parallel or antiparallel
alignment would not lead to torques). If the electron flow is opposite (Fig. 24b),
the electrons will first undergo spin filtering by the free layer. They will then flow to
the fixed layer and apply a torque to that layer. This torque, however, is inefficient
as the fixed layer magnetization is held rigidly in place. However, a fraction of
the electrons will be reflected back from the interface between normal metal and
fixed layer toward the free layer. These reflected electrons have an averaged spin
polarization antiparallel to the fixed layer magnetization (the parallel spins are
readily transmitted), i.e., opposite to the case discussed before. The free layer now
feels a torque that turns the free layer moment away from the fixed layer moment,
thus destabilizing the parallel alignment of magnetization and possibly leading to
a reversal of the free layer magnetization if the current is sufficiently strong. The
fixed layer consequently serves as an electron polarizer by providing the polarized
electrons that finally act on the free layer for both current directions. This current-
induced magnetization switching between parallel and antiparallel magnetization
of free and fixed layer provides a smart alternative to magnetization switching by
magnetic field and is applied in the STT-MRAM for the writing process.
The dynamics of the free layer in the spin-valve pillar is phenomenologically
described by the LLGS equation (16), in which the τ -term reads

dmfree I
τ= = g(θ ) mfree × (mfree × mfixed ). (25)
dt A

The Slonczewski term, g(θ ), represents the material-dependent spin-transfer effi-


ciency with θ being the angle between the magnetization directions of free and fixed
layers. The associated torques are illustrated in Fig. 24c. For simplicity, a uniaxial
anisotropy along the z-axis is assumed (defining Heff ) with the magnetization of
the fixed layer pointing along the same direction. In the absence of STT, the
magnetization of the free layer, instantaneously pointing at an angle with respect
to z, spirals toward the anisotropy axis, driven by the precessional and damping
torques (compare Fig. 20). In the presence of STT, the additional Slonczewski
torque (25) can point either in the same direction as the damping torque or opposite
to it, depending on the direction of current (compare Figs. 20 and 24c). So the STT
can either reinforce the damping torque, making the free layer to relax toward the
easy axis faster than without, or it can act against the damping so that free layer
magnetization relaxes more slowly. If, in the latter case, the current density exceeds
a critical value, the STT can be larger in magnitude than the damping torque, so
that the free layer begins to spiral away from the z-axis rather than relaxing toward
this direction. If the energy, added to the free layer by the STT, is high enough,
the free layer will finally spiral to θ = π , where it can remain in a stable state
antiparallel to the fixed layer. Depending on the detailed angular dependence of spin
transfer and damping torques, it is also possible that the energies lost to damping and
432 R. Schäfer

gained from the STT are balanced over each precessional cycle. Some dynamical
equilibrium may then be reached at an intermediate angle θ , meaning that steady-
state precessional oscillations in the microwave frequency range are generated [45].
The precession, which is basically caused by a DC applied current, occurs with large
precession angles that are not accessible using magnetic field excitation alone and
which are the basis for the so-called spin-transfer oscillators that are envisaged for
applications in communication technology.

References
1. Hubert, A., Schäfer, R.: Magnetic Domains. The Analysis of Magnetic Microstructures.
Springer, Berlin/Heidelberg (1998)
2. Kirilyuk, A., Ferré, J., Grolier, V., Jamet, J.P., Renard, D.: Magnetization reversal in ultrathin
ferromagnetic films with perpendicular anisotropy. J. Magn. Magn. Mat. 171, 45–63 (1997)
3. Ma, L., Gilbert, D.A., Neu, V., Schäfer, R., Zheng, J.G., Yan, X.Q., Shi, Z., Liu, K., Zhou,
S.M.: Magnetization reversal in perpendicularly magnetized L10 FePd/FePt heterostructures.
J. Appl. Phys. 116(3), 033922 (2014)
4. He, P., Ma, X., Zhang, J.W., Zhao, H.B., Lüpke, G., Shi, Z., Zhou, S.M.: Quadratic scaling of
intrinsic Gilbert damping with spin-orbital coupling in L10 FePdPt films: experiments and Ab
Initio Calculations. Phys. Rev. Lett. 110, 077203 (2013)
5. Lopatina, E., Soldatov, I., Budinsky, V., Marsilius, M., Schultz, L., Herzer, G., Schäfer, R.:
Surface crystallization and magnetic properties of Fe84.3 Cu0.7 Si4 B8 P3 soft magnetic ribbons.
Acta Mat. 96, 10–17 (2015)
6. McCord, J., Schäfer, R.: Domain wall asymmetries in Ni81 Fe19 /NiO: proof of variable
anisotropies in exchange bias systems. New J. Phys. 11, 83016 (2009)
7. Flohrer, S., Schäfer, R., McCord, J., Roth, S., Schultz, L., Herzer, G.: Magnetization loss and
domain refinement in nanocrystalline tape wound cores. Acta Mat. 54 (12) 3253–3259 (2006)
8. Parkin, S.S.P., Hayashi, M., Thomas, L.: Magnetic domain-wall racetrack memory. Science
320, 190–194 (2008)
9. Landau, L.D., Lifshitz, E.: On the theory of the dispersion of magnetic permeability in
ferromagnetic bodies. Phys. Z. Sowjetunion 8, 153–169 (1935)
10. Dzyaloshinskii, I.E.: Theory of helicoidal structures in antiferromagnets. I. Nonmetals. Sov.
Phys. JETP 19, 960–971 (1964)
11. Bogdanov, A.N., Yablonsky, D.A.: Thermodynamically stable "vortices" in magnetically
ordered crystals. The mixed state of magnets. Zh. Eksp. Teor. Fiz 95(1), 178–182 (1989)
12. Bogdanov, A.N., Hubert, A.: Thermodynamically stable magnetic vortex states in magnetic
crystals. J. Magn. Magn. Mater. 138, 255–269 (1994)
13. Yu, X.Z., Onose, Y., Kanazawa, N., Park, J.H., Han, J.H., Matsui, Y., Nagaosa, N., Tokura, Y.:
Real-space observation of a two-dimensional skyrmion crystal. Nature (London) 465, 901–904
(2010)
14. Romming, N., Hanneken, C., Menzel, M., Bickel, J.E., Wolter, B., von Bergmann, K.,
Kubetzka, A., Wiesendanger, R.: Writing and deleting single magnetic skyrmions. Science
341, 636–639 (2013)
15. Bode, M., Heide, M., von Bergmann, K., Ferriani, P., Heinze, S., Bihlmayer, G., Kubetzka,
A., Pietzsch, O., Blügel, S., Wiesendanger, R.: Chiral magnetic order at surfaces driven by
inversion asymmetry, Nature 447, 190–193 (2007)
16. Leonov, A.O., Monchesky, T.L., Romming, N., Kubetzka, A., Bogdanov, A.N., Wiesendanger,
R.: The properties of isolated chiral skyrmions in thin magnetic films. New J. Phys. 18, 065003
(2016)
17. Mühlbauer, S., Binz, B., Jonietz, F., Pfleiderer, C., Rosch, A., Neubauer, A., Georgii, R., Böni,
P.: Skyrmion lattice in a chiral magnet. Science 323, 915–919 (2009)
8 Magnetic Domains 433

18. Kiselev, N.S., Bogdanov, A.N., Schäfer, R., Rössler, U.K.: Chiral skyrmions in thin magnetic
films: new objects for magnetic storage technologies? J. Phys. D: Appl. Phys. 44, 392001
(2011)
19. Bogdanov, A.: Private communication
20. Hilzinger, R., Rodewald, W.: Magnetic Materials. Publicis Publishing, Erlangen (2013)
21. Herzer, G.: Private communication
22. Soldatov, I., Schäfer, R.: Selective sensitivity in Kerr microscopy. Rev. Sci. Instrum. 88, 073701
(2017)
23. Schäfer, R.: The Magnetic Microstructure of Nanostructured Materials. In: Liu, J., Fullerton,
E., Gutfleisch, O., Sellmyer, D. (eds) Nanoscale Magnetic Materials and Applications.
Springer, Boston (2009)
24. Malozemoff, A.P., Slonczewski, J.C.: Magnetic Domain Walls in Bubble Materials. Academic
Press, New York (1979)
25. Döring, L., Hengst, C., Otto, F., Schäfer, R.: Interacting tails of asymmetric domain walls:
theory and experiments. Phys. Rev. B 93, 024414 (2016)
26. Thiaville, A., Rohart, S., Jué, É., Cros, V., Fert, A.: Dynamics of Dzyaloshinskii domain walls
in ultrathin magnetic films. Europhys. Lett. 100 (5), 57002 (2012)
27. Thiaville, A., Nakatani, Y.: In: Shinjo, T. (eds.) Nanomagnetism and spintronics, p. 231.
Elsevier (2009)
28. Thiaville, A., Nakatani, Y.: In: Hillebrands, B., Thiaville, A. (eds.) Spin dynamics in confined
magnetic structures III, p. 161. Springer, Berlin (2006)
29. Mougin, A., Cormier, M., Adam, J.P., Metaxas, P.J., Ferré, J.: Domain wall mobility, stability
and Walker breakdown in magnetic nanowires. Europhys. Lett. 78 (5), 57007 (2007)
30. Hubert, A.: Theorie der Domänenwände in geordneten Medien. Springer, Berlin/Heidel-
berg/New York (1974)
31. Filippov, B.N.: Static properties and nonlinear dynamics of domain walls with a vortexlike
internal structure in magnetic films (Review). Low Temp. Phys. 28, 707–738 (2002)
32. Berger, L.: Exchange interaction between ferromagnetic domain wall and electric current in
very thin metallic films. J. Appl. Phys. 55, 1954–1956 (1984)
33. Berger, L.: Emission of spin waves by a magnetic multilayer traversed by a current. Phys. Rev.
B 54, 9353 (1996)
34. Slonczewski, J.C.: Current-driven excitation of magnetic multilayers. J. Magn. Magn. Mat.
159, L1–L7 (1996)
35. Thiaville, A., Nakatani, Y., Miltat, J., Suzuki, Y.: Micromagnetic understanding of current-
driven domain wall motion in patterned nanowires. Europhys. Lett. 69, 990–996 (2005)
36. Zhang, S., Li, Z.: Roles of nonequilibrium conduction electrons on the magnetization dynamics
of ferromagnets. Phys. Rev. Lett. 93, 127204 (2004)
37. Thanks to T. Moore (Leeds) and J. Augusto (IFW) for straightening the vectors
38. Parkin, S., Yang, S.-H.: Memory on the racetrack. Nat. Nanotechnol. 10, 195 (2015)
39. Khvalkovskiy, A.V., Cros, V., Apalkov, D., Nikitin, V., Krounbi, M., Zvezdin, K.A., Anane, A.,
Grollier, J., Fert, A.: Matching domain-wall configuration and spin-orbit torques for efficient
domain-wall motion. Phys. Rev. B 87, 020402 (2013)
40. Manchon, A., Koo, H.C., Nitta, J., Frolov, S.M., Duine, R.A.: New perspectives for Rashba
spin-orbit coupling. Nat. Mater. 14, 871–882 (2015)
41. Miron, I.M., Moore, T., Szambolics, H., Buda-Prejbeanu, L.D., Auffre, S., Rodmacq, B.,
Pizzini, S., Vogel, J., Bonfim, M., Schuhl, A., Gaudin, G.: Fast current-induced domain-wall
motion controlled by the Rashba effect. Nat. Mater. 10, 419–423 (2011)
42. Yang, S.-H., Ryu, K.-S., Parkin, S.: Domain-wall velocities of up to 750 m s−1 driven by
exchange-coupling torque in synthetic antiferromagnets. Nat. Nanotechnol. 10, 221–226
(2015)
43. Ralph, D.C., Buhrman, R.: In: Maekawa, S. (ed.) Concepts in spin electronics. Oxford
University Press, Oxford (2006)
44. Stiles, M.D., Miltat, J.: In: Hillebrands, B., Thiaville, A. (eds.) Spin dynamics in confined
magnetic structures III. Springer-Verlag, Berlin, Heidelberg (2006)
434 R. Schäfer

45. Kiselev, S.I., Sankey, J.C., Krivorotov, I.N., Emley, N.C., Schoelkopf, R.J., Buhrman, R.A.,
Ralph, D.C.: Microwave oscillations of a nanomagnet driven by a spin-polarized current.
Nature 425, 380–383 (2003)

Rudolf Schäfer studied Materials Science and received his Ph.D.


in Engineering at Erlangen-University in 1990. After Postdoc
stays at IBM Research in Yorktown Heights and Forschungszen-
trum Juelich he moved to IFW Dresden in 1993. His interest areas
span magnetic materials with focus on magnetic microstructures.
Magnetotransport
9
Michael Ziese

Contents
Basics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 436
Classical Magnetoresistance in Semiconductors and Semimetals . . . . . . . . . . . . . . . . . . . . . . 440
Magnetotransport and Ferromagnetism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 442
Anisotropic Magnetoresistance, Planar Hall Effect, and Two-Current Model . . . . . . . . . . . 442
Giant Magnetoresistance (GMR ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 446
Colossal Magnetoresistance (CMR ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 448
Tunneling Magnetoresistance (TMR ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 451
Powder Magnetoresistance (PMR ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 454
Organic Magnetoresistance (OMAR ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 454
Quantum Transport . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 454
Exotic Tunneling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 456
Hall Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 458
Spin Currents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 460
Spin Hall Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 461
Inverse Spin Hall Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 462
Quantum Spin Hall Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 462
Magnetoimpedance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 464
Measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 464
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 467

Abstract

In this chapter, magnetotransport phenomena in ferromagnetic materials and


heterostructures are discussed. The survey starts with the definition of the
electrical resistivity and a discussion of magnetoresistance in normal met-

M. Ziese ()
Fakultät für Physik und Geowissenschaften, Universität Leipzig, Leipzig, Germany
e-mail: ziese@physik.uni-leipzig.de

© Springer Nature Switzerland AG 2021 435


J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_9
436 M. Ziese

als, semiconductors, and semimetals. This is followed by a description of


magnetoresistance processes in ferromagnets: from anisotropic magnetoresis-
tance and critical scattering in bulk ferromagnets to giant magnetoresistance
in magnetic multilayers, colossal magnetoresistance in manganites, tunneling
magnetoresistance in trilayers with insulating barrier, and to powder and organic
magnetoresistance. The two-current model features prominently in the analysis
of the majority of these effects. Sections on quantum transport and exotic
tunneling phenomena follow. Then transverse electric phenomena are presented:
the discussion of the normal and anomalous Hall effect leads to the introduction
of spin currents and the spin Hall effects. The chapter concludes with a brief
section on magnetoimpedance and an explanation of measurement techniques.

Basics

When a static electric field E is applied to a material, it might respond with a current
flow – then it is called an electric conductor – or with a static polarization P , in
which case it is called an insulator. Conductors are characterized by Ohm’s law [1]:
When all external parameters (temperature, pressure, magnetic field . . . ) are held
constant, the current density j flowing through a conductor is proportional to the

applied electric field E:
j = σ E . (1)

σ is a tensor of second rank known as the conductivity tensor; it is a material


property with units S/m. The inverse tensor ρ = σ−1 is the resistivity tensor
with units  m. When a magnetic field H is applied, the conductivity tensor obeys
Onsager’s relation [2]:
σij (H ) = σj i (−H ) . (2)

The diagonal components are even under magnetic field inversion; the off-diagonal
components might have an even as well as an odd contribution. In case of isotropic
systems and cubic systems, all three diagonal components are equal and are
simply called conductivity σ . In this chapter, the material systems are restricted
to solids; according to their conductivity values, conductors might be quantitatively
further divided into metals (good conductors) and semiconductors. However, it is
more appropriate to use the temperature coefficient of resistivity (TCR) dρ/dT
to qualitatively distinguish metals (positive TCR) from semiconductors (negative
TCR). More than two thirds of the elements are metals; see Fig. 1.
If contributions from ionic motion are neglected, the current flow is carried
by electrons that are driven through the material by the applied electric field. In
a crystalline material, the electrons are found in Bloch states characterized by a
quantum number k,  the wave vector, or h̄k,
 the crystal momentum. The electronic

energy En (k) is a function of band index n and wave vector k, being periodic in the
9 Magnetotransport 437

Fig. 1 Left panel: Room temperature resistivities of the elements as a function of order number.
Resistivities diverge toward the r.h.s. of the periodic table (C (graphite), Si, Ge, Te). Certain
features recur periodically, such as the resistivity minima of the Cu group elements, the local
maxima of the Ti group elements, as well as a maximum at half-filling of the d-shell (Mn, Tc, Re).
Resistivity values from [3]. Right panel: Resistivity of Fe1−x Nix alloys at various temperatures [4].
The resistivity increase of the alloy is clearly seen, especially at low temperatures. The situation is
further complicated by the martensite-austenite transition that occurs along the alloying route from
iron (bcc) to nickel (fcc)

reciprocal lattice and thus forming energy bands [5]. The electron velocity vn (k) 
does in general not vanish, i.e., electrons move through periodic lattices without
scattering. Electron scattering is due to imperfections in the crystal lattice, either by
static imperfections (lattice defects) or by dynamic imperfections such as phonons
or magnons. Within the Drude model , the scattering is characterized by a relaxation
time τ . The conductivity can then be expressed by the carrier density n, the effective
electron mass m∗ , the relaxation time τ , and the electron charge e as [6]

e2 nτ
σ = . (3)
m∗

When various scattering mechanisms act in the same conductor, in a first approx-
imation, the resistivities add up: ρ = ρi + ρp + . . . (Matthiessen’s rule, [7]). ρi
denotes the scattering by impurities that is mainly temperature independent, and ρp
denotes the scattering by phonons that is proportional to T 5 (Bloch T 5 law) at low
temperatures, to T 3 for umklapp processes, and to T at high temperatures [5]. The
effect of electron-electron scattering is rather small leading to a term proportional
to T 2 . As a simple rule, the resistivity of the non-ferromagnetic elements at room
temperature is proportional to the temperature, since electron-phonon scattering
dominates, whereas alloys have a nearly temperature-independent resistivity due to
438 M. Ziese

the dominance of impurity scattering. This is illustrated in the right panel of Fig. 1
showing the resistivity of Fe1−x Nix alloys for various temperatures.
A well-established model describing electron transport in the limit of weak
scattering is the semiclassical model of electron dynamics. This builds on the energy
 According to Hamilton’s equations, the velocity is related to
band structure En (k).
the momentum gradient of the Hamiltonian [5]:

 = 1 ∂En /∂ k .
vn (k) (4)

The time derivative of the crystal momentum is given by the Lorentz force [5]:
 
h̄k˙ = −e E(  × B(
 r , t) + vn (k)  r , t) . (5)

This model was very successful in explaining a wealth of phenomena, especially


the distinction between metals, semiconductors, and insulators as systems with or
without overlapping energy bands, the values of the carrier density, the occurrence
of holes as charge carriers, and magnetoresistance effects. In an applied magnetic
induction B,  both the electronic energy and the wave-vector component along B
are constant, i.e., the electrons move on constant energy orbits perpendicular to
the magnetic field. If an electric field E is applied perpendicular to the magnetic
induction, a drift motion with velocity E × B/B  2 is superimposed on the cyclic
motion. In the limit of high fields, i.e., when the cyclotron frequency ωc = eB/m∗
is large compared to the scattering rate τ −1 , the E × B drift dominates and
leads to a Hall current; see section “Hall Effect ”, if the cyclic orbits are closed,
i.e., restricted in space. In this case the magnetoresistance saturates; in the case
of open orbits extending between adjacent Brillouin zones, the magnetoresistance
increases with increasing magnetic induction without bounds. The electronic motion
in a magnetic field leads to a variety of effects (Shubnikov-de Haas effect in case of
magnetoresistance) being periodic in 1/B with period
 
1 2π e
Δ = , (6)
B h̄A(EF )

where A(EF ) denotes the cross section of an extremal Fermi surface perpendicular
to the applied magnetic field. These effects have proven viable tools for Fermi
surface studies [5].
If the mean free path
= vF τ , where vF denotes the Fermi velocity, is of the
order of the lattice spacing a, the Mott-Ioffe-Regel limit
= a is reached and the
metal approaches insulating behaviour [9]. In the limit of strong disorder, in an alloy
or an amorphous material, the electronic wave function may become localized [10],
such that a crossover or transition to an insulating state with variable range hopping-
dominated conductivity occurs [9]. Analyzing the formation of extended wave
functions from the atomic orbitals, Mott [9] introduced the concept of a mobility
edge that separates localized states in the tails of the energy bands from extended
9 Magnetotransport 439

states within the energy band centers. In certain compounds, this limit is actually
accompanied by a resistivity saturation [11,12], but this is not a universal signature,
and the resistivity of other metals just crosses this limit without any features [13].
Electron scattering in the limit of strong disorder is determined by various scattering
length scales, and especially the phase coherence length can be much larger than
the mean free path. This leads to the possibility that an electron is strongly scattered
and retraces its original trajectory coherently, a process known as weak localization
[14]. Weak localization leads to corrections to the resistivity being logarithmic
in temperature in two-dimensional and proportional to T 1/2 in three-dimensional
systems. Furthermore, a magnetic field threading the weakly localized trajectories
destroys the phase coherence and leads to a negative magnetoresistance [14].
Spin-orbit coupling has a similar effect but leads to a positive magnetoresistance
contribution, see, e.g., the magnetoresistance of Au in Mg [15].
After discussing some of the basics of magnetotransport in metals, one might
ask what the specifics of magnetotransport in ferromagnets are. This is nicely
highlighted by a set of classic data from Gerlach and Schneiderhan [8]; see Fig. 2.
In ferromagnets, electron-magnon scattering is large and contributes another term

Fig. 2 Resistance and magnetoresistance of a nickel wire; magnetic field and current density were
applied along the wire axis. (a) The resistivity of ferromagnetic metals has a slope change at
the Curie temperature (615 K in this case, 633 K for pure Ni). (b) The magnetoresistance slope,
and therefore also the magnetoresistance in large magnetic fields has a maximum at the Curie
temperature due to critical scattering. (c) The magnetoresistance below the Curie temperature
shows an anisotropic contribution at low fields and a linear contribution from critical scattering
at high fields. (d) Above the Curie temperature, only the linear MR remains. The numbers indicate
the measurement temperatures. (After [8])
440 M. Ziese

ρm ∝ T 3/2 to Matthiesen’s rule [16]; see also section “Magnetotransport and


Ferromagnetism”. Above the Curie temperature, magnon scattering diminishes, thus
often leading to a slope change in the resistivity of ferromagnets; see Fig. 2a. The
magnetoresistance measured as a function of magnetic field shows two character-
istics: (1) below the Curie temperature TC , see Fig. 2c; the magnetoresistance rises
sharply in small fields and decreases linearly in higher field; and (2) above TC , only
the linear magnetoresistance remains having a maximum slope at the Curie temper-
ature. The sharp magnetoresistance rise, known as anisotropic magnetoresistance
(AMR), is due to a reorientation of the magnetization and is further discussed in
section “Magnetotransport and Ferromagnetism”. The linear magnetoresistance is
due to magnon scattering and extends to very high fields [17]. Figure 2b shows the
slope ∂(ΔR/R)/∂(μ0 H ) as a function of temperature; a clear maximum is observed
at the Curie temperature that is due to critical scattering.

Classical Magnetoresistance in Semiconductors and Semimetals

In the simplest approximation, the free electron model, magnetoresistance is absent,


but the electrons show a Hall effect ; see section “Hall Effect ”. Electron motion
is perpendicular to the magnetic field and can be described by a two-dimensional
resistivity or conductivity tensor
   
1 −ωc τ 1 1 ωc τ
ρ=ρ σ = . (7)
ωc τ 1 ρ(1 + ωc2 τ 2 ) −ωc τ 1

ρ denotes the longitudinal resistivity; the Hall constant is RH = ρωc τ/B. Note that
the diagonal components of the conductivity depend on the magnetic field. If more
than one band contributes to the electron transport, the total conductivity tensor
is obtained as the sum over the conductivity tensors of all bands. In case of two
bands, this yields the following expressions for the longitudinal resistivity ρ and
Hall constant R (two-band model, the indices refer to the respective bands):

ρ1 ρ2 (ρ1 + ρ2 ) + (ρ1 R22 + ρ2 R12 )B 2


ρ= (8)
(ρ1 + ρ2 )2 + (R1 + R2 )2 B 2
R1 ρ22 + R2 ρ12 + R1 R2 (R1 + R2 )B 2
R= . (9)
(ρ1 + ρ2 )2 + (R1 + R2 )2 B 2

These expressions highlight certain features of the classical magnetoresistance.


In small magnetic fields the resistivity is proportional to B 2 , whereas it sat-
urates in large fields. The Hall constant is given by a weighted average in
small fields and saturates to R → 1/[e(±n1 ± n2 )] in high fields, where the
±-sign indicates either hole or electron conduction. We further see that the strength
of the magnetoresistance is determined by ωc τ . Since the relaxation time τ is
inversely proportional to the zero field resistivity, a scaling of the magnetoresistance
9 Magnetotransport 441

with B/ρ is expected, which is known as Kohler’s rule [20]. A nonsaturating magne-
toresistance is obtained in compensated materials with n1 +n2 = 0. These equations
form the basis for a magnetotransport analysis of semiconductors. They were also
used to describe the behavior of compensated metals such as Bismuth or zero-gap
semiconductors such as graphite. In graphite and Bi, the magnetoresistance is so
strong that a magnetic field-induced metal-insulator transition occurs, i.e., the resis-
tivity at low temperatures is larger than at high temperatures, since ωc τ is very large
at low temperatures; see Fig. 3. The resistivity and magnetoresistance can be fairly
well understood within a three-band model; see Fig. 3a and b [18]. This is somewhat
surprising, since the magnetoresistance in general depends on the form of the Fermi
surface as well as the dominant scattering mechanism [21]. Further, the semiclassi-
cal model is only an approximation and does not provide a full quantum-mechanical
description of the magnetotransport. The observation of a magnetoresistance linear
in magnetic field, e.g., in Ag2+δ Se [22] and in graphite [23] or multilayer
graphene [24], calls for a full quantum-mechanical description [25]. Early quantum
theories of galvanomagnetic effects found a scaling of the magnetoresistance in the
quantum limit with H p /T q , where the exponents p and q depend on the scattering

Fig. 3 In-plane resistivity of graphite as a function of (a) temperature at various fields up to


200 mT and (b) magnetic field at various temperatures up to 200 K. The symbols are data and the
solid lines fit to a three-band model (After [18]). (c) Kohler plot for the in-plane magnetoresistance
of graphite in magnetic fields up to 200 mT. Obviously, Kohler’s rule is violated in graphite. (d)
The same magnetoresistance data as in (c) but this time plotted as a function of H /T . Scaling is
found for temperatures between 20 K and at least 270 K (After [19])
442 M. Ziese

mechanism [26, 27]. Although the problem of quantum magnetoresistance is far


from solved, it is clear that the magnetoresistance of graphite does not obey Kohler’s
rule but comes close to a H /T -scaling; see Fig. 3c and d.

Magnetotransport and Ferromagnetism

Anisotropic Magnetoresistance, Planar Hall Effect, and Two-Current


Model

Spin disorder scattering near the Curie temperature was discussed in sec-
tion “Basics”. Below the Curie temperature, the magnetoresistance of a ferromagnet
depends on the relative direction between the electric current density j and the
magnetization M  [28, 16, 29]. This constitutes the anisotropic magnetoresistance
(AMR ). Let ρ and ρ⊥ denote the resistivities (extrapolated to zero induction B)
in longitudinal (j  M)
 and transverse geometry (j ⊥ M),  respectively; then the
anisotropic magnetoresistance is defined as

Δρ ρ − ρ⊥
= . (10)
ρ 1
3 ρ + 23 ρ⊥

In case of a polycrystalline or amorphous ferromagnet, the resistivity depends only


on the angle Θ between magnetization and current density, but not on the crystal
directions; this gives rise to an angle-dependent longitudinal resistivity

ρlong = ρ⊥ + (ρ − ρ⊥ ) cos2 Θ (11)

and a transverse resistivity

ρtrans = −(ρ − ρ⊥ ) sin Θ cos Θ . (12)

In case of single crystalline materials, the angle dependence might be much more
complicated [29, 30, 31, 32]. As an illustration, Fig. 4 shows the anisotropic magne-
toresistance of a single crystalline SrRuO3 film as a function of angle between the
applied field of 8 T and the crystallographic axes. The complex angle dependence
is obvious. Since the magnetic anisotropy of SrRuO3 is rather strong, the angular
dependence is even further complicated, since the magnetization is not always along
the magnetic field direction; note especially the sharp magnetoresistance change at
about 30◦ from the [110] axis in the (001) plane, which is due to the crossing of a
hard magnetic axis [33].
The transverse resistivity is known as planar Hall effect . Due to the absence
of any offset in Eq. (12), this is sometimes advocated to be a much more sensitive
probe than the AMR. Especially in case of ferromagnets with rather large resistivity,
9 Magnetotransport 443

Fig. 4 Angle dependence of the AMR in a single crystal SrRuO3 film. The current density was
along the [001] orthorhombic axis; the magnetic field was rotated in the (a) (110) and (b) (001)
plane. Some crystallographic directions are indicated (After [33])

the planar Hall effect was dubbed giant [34, 35]. The planar Hall effect of
La0.84 Sr0.16 MnO3 is shown in Fig. 5.
The basis for the understanding of the AMR is Mott’s two-current model
[36, 37]. This in turn is based on the s − d model: the band structure of a 3d
transition metal might be understood as being composed of a broad free-electron-
like s-band superimposed on a narrow tight-binding-like d-band. The comparatively
high resistivity of the transition metals might be understood by scattering of mobile
s-electrons into d-states, where the latter have a considerable density of states at
the Fermi level. Spin-flip scattering events are rare, at least at low temperature,
since scattering occurs predominantly at non-magnetic scattering centers. Therefore,
electron transport can be viewed as a parallel circuit of spin-up and spin-down
conduction channels; see Fig. 6; in each channel, the resistivities ρ ↑ and ρ ↓ consist
in general of a small ss- and a large sd-scattering contribution. The spin-flip-
−1
scattering rate τ↑↓ is negligible at low temperatures. In case of strong ferromagnets,
the majority (or spin-up band) is completely filled, thus eliminating the sd-scattering
of spin-up electrons and short-circuiting carrier transport along the spin-up channel.
Historically, this model was motivated by a resistivity drop below the Curie
temperature; see Fig. 6c; note, however, data and discussion in [16].
For a measurement of ρ↓ and ρ↑ , the scattering center has to be defined. This
is usually done by alloying a ferromagnetic host metal with a few atomic percent
of particular metal impurities. The channel resistivities can then be obtained from
deviations of the measured data from Matthiessen’s rule either as a function of
temperature or in ternary alloys [16]. Figure 7 shows the channel resistivities as well
444 M. Ziese

Fig. 5 (a) AMR and (b) and (c) planar Hall effect of La0.84 Sr0.16 MnO3 at 120 K. The lines in (a)
and (b) are fits of Eqs. (11) and (12) to the data using (a) R − R⊥ = −29 Ω and (b) R − R⊥ =
−88 Ω. (After [35])

Fig. 6 Mott’s two-current model: (a) spin-up and spin-down conduction channels weakly coupled
by spin-flip scattering. (b) Short-circuiting by the spin-up channel in case of a strong ferromagnet.
(c) Similarity of the temperature dependence of the resistivities of Ni (above TC ), Pd, and Pt; for
comparison Ag
9 Magnetotransport 445

Fig. 7 (a) Channel resistivities ρ↓ (solid symbols) and ρ↑ (×ed symbols) of Ni, Co and Fe for the
specified 3d transition metal impurities. (b) Channel resistivity ratio α = ρ↓ /ρ↑ as a function of
impurity

Table 1 Channel resistivities ρ↓ and ρ↑ and α = ρ↓ /ρ↑ for Ni, Co, and Fe hosts. (From [16, 38,
39, 40, 41, 42])
ρ↓ ρ↑ α
(nΩm) (nΩm)
Impurity/host Ni Co Fe Ni Co Fe Ni Co Fe
Ti 105 110 40 56 76 105 1.9 1.4 0.4
V 67 77 13 130 77 105 0.5 1.0 0.1
Cr 65 24 28 220 73 125 0.3 0.3 0.2
Mn 75 100 17 7.5 120 130 10 0.8 0.1
Fe 60 67 4 5.4 15 12
Co 35 33 1.8 16 19 2
Ni 120 24 5
Rh 21 28 64 100 28 11 0.1 1.0 6
Re 75 77 27 260 180 87 0.3 0.4 0.3
Ir 48 38 200 280 117 22 0.2 0.3 9
Os 64 68 43 500 235 130 0.1 0.3 0.3

as their ratio α = ρ↓ /ρ↑ for the ferromagnetic hosts Ni, Co, and Fe as a function of
the 3d transition metal impurities. Table 1 contains data of the channel resistivities.
The data in Fig. 7a conform to the expectation that ρ↑ is small only for transition
metal impurities Fe, Co, and Ni. In case of the early 3d elements, the channel
resistivities are sizable, and for V and Cr that form virtual bound states at the Fermi
level of the ferromagnetic host, α is even smaller than unity. In case of Ti, rather
446 M. Ziese

large channel resistivities are found, but α, at least for the hosts Ni and Co, is larger
than unity.
The AMR is a spin-orbit interaction effect. It might be described by a model
Hamiltonian that takes the exchange interaction, the crystal-field splitting, and the
spin-orbit coupling into account [43, 44, 45]:

 · S .
Ĥ = Ĥex + ĤCF + AL (13)

Spin-orbit coupling leads to a spin-admixture such that the majority carrier wave
functions acquire a small spin-down component and vice versa. The scattering
matrix element due to a spherically symmetric scatterer from an s- into a d-state
is dependent on the magnetization, i.e., spin direction, thus leading to the AMR. In
second-order perturbation theory, one obtains

Δρ
= γ (α − 1) , (14)
ρ

where α = ρ↓ /ρ↑ is the resistivity ratio between up- and down-channels and γ
sets the strength of the effect. In second-order perturbation theory γ ∝ (A/Eex )2
or γ ∝ (A/ECF )2 , depending on whether the exchange splitting Eex or crystal
field splitting ECF is larger. Figure 8 shows the anisotropic magnetoresistance as
a function of channel resistivity ratio α. Up to α 3, the data fall onto a straight
line with a slope of 0.03; shown is a linear dependence with a slope γ = 0.01 as
suggested in [46].

Giant Magnetoresistance (GMR )

Giant magnetoresistance is a phenomenon found in metallic multilayers [47, 48,


49] and spin valves [50]. It occurs when the ferromagnetic layers within the
heterostructure are antiferromagnetically coupled through the adjacent metallic
layers; in a sufficiently large applied magnetic field, the ferromagnetic layers
will be ferromagnetically aligned, and the GMR ratio appears as the normalized
resistance difference between the antiparallel and parallel configuration. Figure 9
shows extensive magnetization and magnetoresistance data on Co/Ru multilayers
that show both the appearance of an antiferromagnetic coupling between the Co
layers that oscillates with the Ru layer thickness as well as the accompanying giant
magnetoresistance showing the same oscillations.
In a first approximation, GMR might be understood within a simple resistor
model [52, 53, 54]. One distinguishes two configurations with the current density
parallel to the layers (current-in-plane, CIP) and perpendicular to the layers
(current-perpendicular-to-plane, CPP); see Fig. 10a; in general CPP GMR is larger
than CIP GMR. Within the spirit of the two-current model, the current density is
9 Magnetotransport 447

Fig. 8 Anisotropic magnetoresistance AMR for alloys of Ni as a function of the channel resistivity
ratio of the particular alloying element. The red line was calculated from Eq. (14) using γ = 0.01
(After [46] and [16])

carried by spin-up and spin-down currents within each layer. A multilayer can then
be modeled by a parallel circuit of layer resistances in series (CPP) or partially in
series (within a mean free path, CIP); see Fig. 10b and [52, 53, 54].
On a more fundamental basis, spin accumulation at the interfaces has to
be taken into account. The concept of spin diffusion was studied in early
work [55] with measurements of the spin diffusion length λs in Al after spin
injection from permalloy. At the interface between a ferromagnet and a normal
metal, the conversion of an unpolarized current in the normal metal into spin-
polarized currents in the ferromagnet leads to a difference in the electrochemical
potentials μ↑ and μ↓ for spin-up and spin-down electrons; see Fig. 10c; this
potential difference in turn causes an additional boundary resistance [56].
Using this concept, a quantitative theory of the GMR in CPP configuration was
developed [57].
In antiferromagnetic spintronics, the spin valve as a trilayer formed by two
ferromagnets separated by a normal metal is replaced by a ferromagnet adjacent
to an antiferromagnetic tunnel junction, e.g., in a NiFe/IrMn/MgO/Pt stack [58].
The resistance change is due to tunneling anisotropic magnetoresistance [59]. The
coupling to the external field is facilitated by the permalloy layer that rotates the
antiferromagnetic magnetic moments by the exchange-spring effect.
448 M. Ziese

Fig. 9 (a)–(f) Magnetization at 300 K (left axis) and transverse magnetoresistance at 4.5 K (right
axis) of six Ru/Co superlattices with structure Si(111)/(10 nm) Ru/[(1.6 nm) Co/tRu Ru]20 /(5 nm)
Ru and Ru layer thicknesses of 0.8, 1.1, 1.9, 2.7, 3.1, and 3.8 nm. (g) Transverse magnetoresistance
and (h) saturation magnetic field as a function of Ru layer thickness. Clear oscillations in
the saturation field and the magnetoresistance are observed: when the interlayer coupling is
antiferromagnetic with a large saturation field, the magnetoresistance is “giant” (After [51])

Colossal Magnetoresistance (CMR )

The isotropic magnetoresistance near the Curie temperature of ferromagnetic oxides


of the type La0.7 Sr0.3 MnO3 (manganites) is called colossal magnetoresistance
[60, 61, 62]. In principle, this magnetoresistance phenomenon is similar to the
critical scattering in metallic ferromagnets; compare Fig. 2. The effect, however, is
significantly larger, especially for compounds with a rather low Curie temperature
[63]. A phenomenological overview of magnetoresistance data is given in Fig. 11a
and b; the magnetoresistance was defined by the pessimistic definition Δρ/ρ0 =
[ρ0 − ρ(B)] /ρ0 , where ρ0 denotes the resistivity value in zero field.
Manganites crystallize in the perovskite structure (RE)1−x Ax MnO3 with rare
earth elements RE = La, Nd, Pr, Y, . . . and A = Ca, Sr, Ba, Pb, or alkaline
elements [68, 69]. The phase diagram of these materials is rather complex; see
Fig. 11c: depending on the doping concentration x, a variety of antiferromagnetic,
ferromagnetic, charge, or orbitally ordered phases is observed [67]. Depending on
the size of the ions, a ferromagnetic phase is found around a doping level of 30%;
maximum Curie temperature is 360 K for Sr-doping.
9 Magnetotransport

Fig. 10 (a) Current-in-plane (CIP) and current-perpendicular-to-plane (CPP) configurations. (b) Resistor model of GMR – upper row: carrier scattering in the
ferromagnetic and antiferromagnetic configurations (schematic); lower row, corresponding spin-dependent circuit models. (c) Spin-dependent electrochemical
potentials at a ferromagnet-normal metal interface; λsF and λsN denote the spin-diffusion lengths in the ferromagnet and normal metal ((a) and (b) after [54])
449
450 M. Ziese

Fig. 11 (a) Zero field resistivity (left axis) and magnetization (right axis) as well as magne-
toresistance ratio in a magnetic field of 1 T of La0.7 Ba0.3 MnO3 and La0.7 Ca0.3 MnO3 films.
(After [64]). (b) Magnetoresistance ratio of various manganite compounds as a function of
the Curie temperature. (After [65] and [63]). (c) Phase diagrams of La1−x Cax MnO3 and
Pr1−x Cax MnO3 with the following phases: PMI paramagnetic insulator, FMM ferromagnetic
metal, FMI ferromagnetic insulator, AFMI antiferromagnetic insulator, CO charge ordered. TC ,
TN , TP , and TCO denote the transition temperatures into the magnetically ordered and charge
ordered phases, TS denotes the temperature of the structural transition between orthorhombic and
rhombohedral phases. (After [66] and [67])

In a first approximation, the electronic structure might be understood from the


octahedral crystal field around the Mn ions. This leads to an energetic splitting of the
threefold degenerate t2g states (dxy , dyz , and dzx orbitals) and the twofold degenerate
eg states (dx 2 −y 2 and d3z2 −r 2 orbitals); see Fig. 12a. The doping with divalent atoms
on the rare earth site introduces a mixed valence of Mn3+ and Mn4+ ions. The strong
Hund’s rule coupling leads to a core spin of 3/2 Bohr magnetons of the low-lying t2g
levels; the additional electron in the eg levels of a Mn3+ ion can hop to a neighboring
Mn4+ ion, when the core spins are aligned; i.e., the hopping integral depends on
9 Magnetotransport 451

Fig. 12 (a) Crystal-field splitting of the Mn 3d-levels in an octahedral environment. The arrows
indicate the electron occupation of the levels for Mn3+ and Mn4 ions according to Hund’s rules.
The dotted arrows indicate hopping of the eg electron from Mn3+ to Mn4+ via the O 2p orbital.
This leads to the two degenerate Mn-O-Mn clusters shown. (c) Calculated band structure and
density of states of La0.7 Sr0.3 MnO3 . The band structure is shown near the Γ point in the directions
toward the X and K points (After [72])

the angle Θ between the core spins as cos(Θ/2). This conduction mechanism is
called double exchange [70,71]; see Fig. 12a; it explains the strong interdependence
between ferromagnetism and conductivity and forms the basis for understanding
colossal magnetoresistance.
The physics of manganites is really intricate, since not only charge and spin
degrees of freedom are coupled, but there is also a coupling to the orbitals and
to lattice phonons [73]. Moreover, the inherent disorder created by the doping
has to be taken into account. This has led to a variety of models for the colossal
magnetoresistance, taking electron-phonon [74], electron-polaron [75], or phase
separation [76] into account. At surfaces and interfaces, orbital reconstruction
might play a role and might lead to new effects [77]. The strong Hund’s rule
splitting leads to a half-metallic [78] band structure; see Fig. 12b [72], and in oxide
heterostructures, La0.7 Sr0.3 MnO3 is the established material for ferromagnetic elec-
trodes or spin injectors. The half-metallicity of the manganites in combination with a
breakdown of the double-exchange mechanism at extended lattice defects leads to a
significant extrinsic magnetoresistance known as grain-boundary magnetoresistance
[79, 80, 81].

Tunneling Magnetoresistance (TMR )

Early work on spin-dependent tunneling focused on tunneling from ferromagnets


through a thin insulating barrier into superconductors [82, 83, 84]. The spin-
dependent density of states D↑ (E), D↓ (E) of a ferromagnet can be measured by the
asymmetry of the conductance peaks that occurs when electrons tunnel through an
452 M. Ziese

AlOx barrier into a thin Al film in a large parallel field that splits the quasiparticle
density of states due to the Zeeman energy; see Fig. 13a. This technique enables
measurements of the spin polarization

D ↑ − D↓
P = . (15)
D↑ + D↓

In related experiments with superconducting tips on ferromagnets, the spin polar-


ization was determined by its influence on Andreev reflection [85,86]. Note that the
spin polarization determined in different experiments might not be identical, since
it might be either related to the density of states (spectroscopic spin polarization,
Eq. (15)) or to the current density (P = (j↑ − j↓ )/(j↑ + j↓ ), compare to Mott’s
two-current model above) [87].
Tunneling between two ferromagnets through an insulating barrier was discussed
for the first time in [88], see data on Co/Ge/Co junctions in Fig. 13b and [104]
for a review. In that work, a simple expression for the tunneling magnetoresistance
between two ferromagnets with spin polarizations P1 and P2 was derived:

R↑↓ − R↑↑ 2P1 P2


T MR = = ; (16)
R↑↑ 1 − P1 P2

R↑↑ and R↑↓ denote the resistance values for parallel and antiparallel magnetization
orientation; note that in the standard pessimistic MR definition used in this chapter
T MR = 2P1 P2 /(1 + P1 P2 ). High-quality ferromagnetic tunnel junctions were
reported for the first time in [89, 105]; see Fig. 13c. The junction resistance is
considerably enhanced in the magnetic field range between the coercive fields of
the two layers, when the magnetizations are antiparallel. The tunneling transport
mechanism is evidenced by the so-called Rowell-criteria [106, 90]: (i) exponential
thickness dependence of the resistance, (ii) quasi-parabolic differential conductance
vs. voltage curves, and (iii) insulator-like temperature dependence of the tunnel
resistance; see Fig. 13d. It is even possible to measure the spin-resolved density
of states using voltage-dependent TMR [107]. As seen from Eq. (16), the TMR
ratio diverges in case of half-metallic ferromagnets; see, e.g., data on LSMO [91] in
Fig. 13e. However, in case of oxide ferromagnets, the TMR ratio decreases strongly
even for temperatures far below the Curie temperature [108]. In certain single
crystalline systems, e.g., Fe/MgO/Fe, details of the band structure might enhance
the TMR ratio above the value estimated by Eq. (16) [109, 92, 110, 111]; see the
data in Fig. 13f. Large TMR ratios were also found for amorphous ferromagnets
[112].
A vast number of spin-polarization values can be found in the literature; Table 2
shows a small selection of consolidated values obtained by the transport techniques
discussed here in comparison to results from spin-polarized photoemission spec-
troscopy. The relation of the spin polarization to band structure and magnetic order
might be complex, see, e.g., CoGd alloys with compensation point [113].
9 Magnetotransport 453

Fig. 13 (a) Conductance G of a Ni/AlOx /Al junction as a function of applied voltage. The con-
ductance was normalized to the conductance GN in the normal state. T = 0.4 K. The conductance
peaks split in magnetic fields applied parallel to the Al film. In a magnetic field, the conductance
peak heights are different, since the density of states in Ni is spin dependent. (After [82]). (b)
Relative conductance ΔG/G0 = (G↑↑ − G↑↓ )/G0 of a Co/Ge/Co junction normalized to the
conductance at 0 V. T = 4.2 K. (After [88]). (c) Tunneling magnetoresistance of a CoFe/Al2 O3 /Co
junction at 295 K. (After [89]). (d) Resistance of a IrMn/NiFeCo/CoFe/AlOx /NiFeCo junction
as a function of temperature in the parallel and antiparallel magnetization state. (After [90]).
(e) Tunneling magnetoresistance of two La2/3 Sr1/3 MnO3 /SrTiO3 /La2/3 Sr1/3 MnO3 junctions
showing the strong decrease with temperature. The inset shows a resistance hysteresis loop at
4.2 K. (After [91]). (f) Tunneling magnetoresistance hysteresis loops for a Fe/MgO/Fe junction at
80 K (open symbols) and 293 K (solid symbols). (After [92])

Table 2 Spin polarization of various ferromagnets determined from transport measurements


(second row) and spin-polarized photoemission spectroscopy [93] (third row). LSMO =
La0.7 Sr0.3 MnO3
Fe Co Ni LSMO CrO2 Fe3 O4 SrRuO3
0.45 [94] +0.42 [94] +0.31 [94] 0.95 [91] 0.90 [95] −0.4 [96] −0.55 [97]
0.4 [98] −0.4 [99] −0.3 [100] 0.9 [101] 0.95 [102] −0.55 [103] ———–

Spin-dependent tunneling is used as a surface-sensitive high-resolution magnetic


imaging technique [114]. To this end, ferromagnetic or antiferromagnetic tips are
used in a STM (scanning tunneling microscope) with atomic resolution, e.g., to
measure the sharpness of domain walls [115] and the nanoscale spatial modulations
454 M. Ziese

of the TMR [116] and to investigate the surface spin chirality [117] or single
skyrmions [118].

Powder Magnetoresistance (PMR )

Powder compacts of oxide ferromagnets often show a strong increase in their


magnetoresistance compared to the respective bulk materials [119, 120, 121].
Powder magnetoresistance is related to grain boundary or tunneling magnetoresis-
tance: the contact areas between adjacent grains are poorly conducting such that
carrier transport occurs by tunneling between regions with different magnetization
directions [69, 81].

Organic Magnetoresistance (OMAR )

Polyfluorene films sandwiched between non-magnetic electrodes (Al, indium-tin-


oxide (ITO)) showed a sizable negative magnetoresistance in magnetic fields of
the order of 100 mT that could be controlled by the applied voltage [122]. Even
the sign of the magnetoresistance could be tuned by varying applied voltage and
temperature [123, 124] or by adjusting the electron and hole injection in double-
layer OLEDs [125]. The OMAR is believed to be an intrinsic property of the
organic semiconductor, not a property of the metal/organic semiconductor/metal
sandwich. The magnetic field dependence has a characteristic shape described by
B 2 /(B02 + B 2 ), where B0 is a constant of the order of 10 mT. OMAR mechanisms
involve the external magnetic field manipulation of paramagnetic entities that are
coupled to the local hyperfine field. These entities might be bipolarons, oppositely
charged polarons, or excitons formed therefrom [126]. Ferromagnetic ordering in
polymers showing OMAR was reported, so there might even be a relation between
the negative magnetoresistance and ferromagnetism [127].

Quantum Transport

In previous sections, transport was mainly in the diffusive regime. In case of mean
free paths being large compared to sample dimensions, i.e., in the ballistic regime,
the sample conductance does not diverge but is found to be quantized in units of
G0 = 2e2 / h = 7.74809173 × 10−5 S. The conductance can then be expressed in
the form [128]

2e2 
G= Ti , (17)
h
i
9 Magnetotransport 455

where Ti denotes the transmittivity of the ith conduction channel. The quantization
can be clearly demonstrated in constrictions or nanowires, when only a few
conduction channels are available, see, e.g., [129,130]. Equation (17) is valid, when
the scattering process does not significantly depend on the spin and when states with
different spin degrees of freedom are energetically degenerate. These conditions do
not necessarily hold in ferromagnetic nanocontacts. Indeed, break junctions made
from Ni wires touching a Ni plate showed conduction with half conductance quanta
e2 / h but only in magnetic fields high enough to saturate the Ni wire magnetization
[131]; see Fig. 14a and b. This indicates that the spin structure close to the contact
is crucial in determining the conductance value.
The magnetoresistance of Ni nanocontacts and break junctions showed large
values up to 80% [132, 134]; see Fig. 14c. Although the data scatter strongly,
a clear trend toward large magnetoresistance values at small conductance values
of a few conductance quanta was observed. Similar results were found for break
junctions between various half-metallic oxides [133, 134] (see Fig. 14d) but also
for nanoconstrictions patterned in half-metallic oxide films [135]. The experimental
data on oxides differ from those on metallic ferromagnets in that the conductance
is significantly smaller than one conductance unit, such that the transport cannot
be ballistic. Domain walls in nanoconstrictions can be very sharp [136]. The

Fig. 14 Histograms of junction conductance for a variety of Ni wire-Ni plate break junctions in
applied magnetic fields of (a) 0 mT and (b) 10 mT. When the Ni wire is in magnetic saturation,
the emergence of spin-polarized half conductance quanta e2 / h was observed (After [131]).
Magnetoresistance of (c) Ni and (d) magnetite nanocontacts as a function of junction conductance.
Note that conductance in case of Ni is ballistic, whereas it is by a tunneling mechanism in case of
Fe3 O4 . (c) After [132]. (d) After [133]. The lines are guides to the eye
456 M. Ziese

magnetoresistance of these nanocontacts is modeled by spin scattering at the domain


wall [137, 134, 133]: in case of a sharp wall, the situation is similar to that in spin-
dependent tunneling with the tunneling probability being proportional to the product
of the density of states left and right to the domain wall; for thicker domain walls,
the resulting magnetoresistance is decreased by an adiabatic factor describing the
accommodation of the travelling spin to the local spin in the domain wall [137,134].
The domain wall magnetoresistance increases strongly with decreasing domain wall
width δ which explains the trends in Figs. 14c and d; in case of oxides, a δ −3
dependence was predicted [81].

Exotic Tunneling

Exotic tunneling refers to spin-dependent tunneling in systems with an active


tunneling barrier [138]. In this section, spin blockade, resonant tunneling, spin
filters, and ferromagnetic and multiferroic barriers will be briefly discussed.
Spin blockade effects occur in systems in which tunneling is via ferromagnetic
nanoparticles. Experimentally this might be realized in granular films consisting
of ferromagnetic nanoparticles embedded in an insulating matrix [139] or by inte-
grating ferromagnetic nanoparticles in the insulating barrier between ferromagnetic
electrodes [144] or paramagnetic electrodes [145,146]. If the electrostatic energy of
the charged nanoparticle is sufficiently large to prevent the tunneling of an additional
electron onto the nanoparticle, the systems are in Coulomb-blockade regime. The
tunneling current then has a steplike dependence on the voltage (Coulomb staircase);
this was observed at room temperature for tunneling through Co nanoparticles
with diameters 1–4 nm embedded in Al2 O3 [146]. The spin accumulation on
the ferromagnetic island was theoretically shown to lead to a further chemical
potential shift that in turn yields an additional modulation of the current-voltage
characteristics leading to an oscillating tunneling magnetoresistance [147,148,149].
Experimentally, in granular systems, one finds a low temperature upturn in the TMR
due to higher order tunneling processes over more than one ferromagnetic island
[139, 120]; see Fig. 15a. Further, an oscillating TMR was reported for tunneling
between a Co and an Al layer through a Co-Al-O film; see Fig. 15b.
Resonant tunneling occurs through some kind of resonant structure, e.g., a
quantum well or a double barrier. A classical device is the Esaki diode showing
a negative differential resistance in the forward bias regime of heavily doped Ge
p-n-junctions [150]. A spin-dependent tunneling device with a quantum well is
shown in the inset to Fig. 15c [140]. A Cu quantum well in a Co/AlOx /NiFe
structure leads to the oscillation of the TMR with Cu barrier thickness. This is due to
the formation of two quantum well states with momenta determined by the Cu Fermi
surface that resonate in the well and lead to an efficient injection and extraction into
the tunneling barrier [140, 151].
As already indicated in the discussion of spin-blockade effects, spin-dependent
tunneling can also be observed between a ferromagnetic and non-magnetic elec-
trode, when these are separated by a thin ferro- or ferrimagnetic material that acts
9 Magnetotransport 457

Fig. 15 (a) TMR of Co-Al-O granular films as a function of temperature. (After [139]). (b) TMR
of a thin Co-Al-O layer in CPP geometry as a function of bias voltage. (After [139]). (c) TMR
of a Co/AlOx /NiFe spin-valve with a Cu quantum wall as a function of Cu layer thickness. (After
[140]). (d) TMR of a La0.7 Sr0.3 MnO3 /NiFe2 O4 /Au spin-filter valve. (After [141]). (e) TMR of a
Fe/BaTiO3 /La0.7 Sr0.3 MnO3 tunnel junction for two orientations of the BaTiO3 polarization. (After
[142]). (f) TMR of a La0.7 Sr0.3 MnO3 /La0.1 Bi0.9 MnO3 /Au spin-filter valve with multiferroic spin
filter. (After [143])

as a spin filter. Due to the exchange energy an electron experiences when tunneling
through the magnetically ordered insulating barrier, the barrier heights for the two
spin directions differ by 2Eex [152, 153]. A more detailed analysis shows that not
only the barrier height but also the band alignment with the electrodes plays a
role [154]. Experimentally, spin filtering was observed for ferromagnetic EuO in
Al/EuO/Y/Al junctions by conductance spectroscopy in the superconducting state
of Al [155] and ferrimagnetic NiFe2 O4 and multiferroic La0.1 Bi0.9 MnO3 by mag-
netoresistance measurements in La0.7 Sr0.3 MnO3 /NiFe2 O4 /Au [141] (see Fig. 15d)
and La0.7 Sr0.3 MnO3 /SrTiO3 /La0.1 Bi0.9 MnO3 /Au [156] tunneling junctions.
Another way of modulating the barrier height is by the use of a ferroelectric
barrier with the electric polarization along the tunneling direction. In this case,
the barrier height depends on the direction of polarization [157, 158]. This might
be due to electrostatic effects, to strain, or to detailed cation shifts at the interface
between the ferroelectric and the electrodes [157]. It is often difficult to distinguish
between a barrier height effect or extrinsic mechanisms such as resistance switching
458 M. Ziese

that might depend on interfacial effects or the formation of conducting paths


within the ferroelectric [159, 160]. The use of ferromagnetic electrodes and a
ferroelectric barrier allows for the construction of a four-state memory element. This
has been demonstrated, e.g., for Fe/BaTiO3 /La0.7 Sr0.3 MnO3 [142]; see Fig. 15e,
for La0.7 Sr0.3 MnO3 /PbZr0.2 Ti0.8 O3 /Co [161] and for La0.7 Sr0.3 MnO3 /PbTiO3 /Co
junctions [162].
A four-state logic might also be realized using a multiferroic spin filter. In this
case, the spin filter can be modulated by application of a magnetic field but also by
application of an electric field via a voltage pulse. Experimentally, such a device was
realized for the first time in a La0.7 Sr0.3 MnO3 /La0.1 Bi0.9 MnO3 /Au tunnel junction
[143]; see Fig. 15f. For further reading, the review papers [163, 164, 165] are
recommended.

Hall Effect

When a magnetic induction B is applied perpendicular to a flat sample, then a


transverse electric field Ey compensating the Lorentz force is generated by charge
accumulation at the sample edges. This transverse electric field gives rise to a
transverse voltage VH , called Hall voltage [166]; see Fig. 16a. In a metal, electrons
(−) or holes (+) may be charge carriers, and the normal Hall constant is given by

VH d Ey ρyx 1
RH = = = =± . (18)
Ix B jx B B en

In case two bands are contributing to the carrier transport, the Hall constant was
already given in Eq. (9). The Hall angle is defined as tan ΘH = Ey /Ex = ρyx /ρxx .
A magnetic field applied to a ferromagnetic metal aligns the magnetization along
its direction; see Fig. 16a. The magnetization component M along the magnetic field
direction leads to an additional Hall effect contribution, known as extraordinary or
anomalous Hall effect. Phenomenologically, the Hall resistivity is then given by
[170]

ρyx = RH B + μ0 RA M (19)

with the anomalous Hall constant RA . Note that in ferromagnetic films the demagne-
tizing factor is close to unity and the magnetic field H within the sample is given in
terms of the applied field HA by H = HA − NM, such that the magnetic induction
is B = μ0 (H + M) = μ0 (HA + (1 − N)M) μ0 HA . Figure 16b shows the
Hall resistivity of a La0.7 Sr0.3 MnO3 film as a function of magnetic field. At low
fields, the Hall resistivity rises steeply due to the anomalous Hall contribution (see
Fig. 16c); at high fields the magnetization M is saturated, and a much smaller slope
due to the ordinary Hall effect is observed. The product μ0 RA MS is obtained by
extrapolation to B = 0 as indicated in Fig. 16b; RH is obtained from the high field
slope. The manganites are hole conductors with a negative anomalous Hall constant.
9 Magnetotransport 459

The anomalous Hall effect is due to the spin-orbit interaction; see section “Spin
Currents ”. One distinguishes between extrinsic and intrinsic contributions. Extrin-
sic contributions are due to scattering: (i) Skew-scattering is the asymmetric
scattering of an electron with respect to the plane containing its incident velocity
and the ion’s magnetic moment [170, 171, 172]. (ii) Side-jump scattering denotes
the side shift of the extrapolated trajectories before and after the scattering event
[170, 173]. Both mechanisms predict a scaling RA ∝ ρ n of the anomalous Hall
constant with the longitudinal resistivity, with n = 1 for skew scattering and n = 2
for side-jump scattering, compare to Fig. 16d.
The intrinsic contribution to the anomalous Hall effect is due to the Berry phase
electrons acquire when moving through a spin lattice [174, 175]. An example
material for this contribution might be SrRuO3 which has an intricate temperature
dependence of the anomalous Hall constant attributed to magnetic monopole-like
features at the Weyl points in momentum space [176, 177]; this is illustrated

Fig. 16 (a) Hall effect in a ferromagnet. (b) Hall resistivity ρyx measured on a 5 nm thick
La0.7 Sr0.3 MnO3 film. At higher temperature, the anomalous Hall effect is clearly seen; the
extrapolation to B = 0 to determine μ0 RA MS is indicated. (c) Initial Hall effect slopes dρyx /dB
for 300 nm thick La0.7 Ca0.3 MnO3 and La0.7 Ba0.3 MnO3 films as well as a Ni sample. (After [64]
and [167]). (d) Scaling of the Hall constant with the longitudinal resistivity for the films from
(c). Solid and dashed lines indicate the linear and quadratic scaling laws due to skew- and side-
jump scattering. (After [64]) (e) Hall resistivity of SrRuO3 illustrating the sign change of the
anomalous Hall constant. (After [168]). (f) Hall effect of a SrRuO3 /SrIrO3 bilayer decomposed
into the anomalous (AHE) and topological (THE) Hall effect contributions (After [169])
460 M. Ziese

by Fig. 16e showing the sign change of the anomalous Hall constant in SrRuO3 .
Further, Berry phase effects lead to the emergence of a topological Hall effect when
the charge carriers move through spin structures with complex topology such as
skyrmions [178,169]. Although the spin topology does not leave a direct signature in
the magnetization component M, an additional Hall resistivity ρT H E , not following
the field dependence of the magnetization, is observed. Complex spin structures
might be induced by interfacial spin-orbit coupling, as illustrated in Fig. 16f showing
the Hall effect of a SrRuO3 /SrIrO3 bilayer [169].
The Hall effect can be used for nanoscale magnetometry, since Hall bars can be
fabricated on the micro- and nanometer scale; see, e.g., magnetometry on single iron
nanoparticles [179] or the discovery of flux-line lattice melting in Bi2 Sr2 CaCu2 O8
high temperature superconductors [180].

Spin Currents

The charge current density from Eq. (1) is defined as j = (−e)n v with the electronic
charge (−e), the carrier density n, and the electronic velocity v. Taking not only
drift, but also diffusion currents into account, Eq. (1) is generalized to

 
∂n 1 ∂μc
ji = σ Ei + eD = σ Ei + (20)
∂xi e ∂xi

with the electronic diffusion constant D. In terms of a generalized force, the


diffusion is not driven by a density gradient, but by the gradient of the chemical
potential μc ; the second equation on the right hand side follows from the generalized
Einstein equation eD(∂n/∂μc ) = nμ, where μ = σ/(en) is the electronic mobility.
In case of a semiconductor, in the non-degenerate case (∂n/∂μc ) = n/(kB T )
which leads back to the Einstein equation. A similar phenomenology holds for spin-
dependent charge currents [181] but is not further followed here.
Instead, the following discussion uses a phenomenological approach proposed
in [182]. Besides the charge the electron also transports spin. The average spin
direction is characterized by the spin polarization vector P , −1 ≤ Pi ≤ 1, such that
the spin density is given by nP . The transport of the spin direction is described by
a second rank tensor [182, 183]


js,ik = nvi Pk . (21)
2

The units of the spin current density js,ik are chosen such that it carries angular
momentum h̄/2. Since the spin transport is coupled to the charge transport, the
spin current density is also driven by electric fields and chemical potential gradients
[182, 183]:
9 Magnetotransport 461

      
h̄ ∂(nPk ) h̄ 1 ∂μs,k
js,ik =− σ Ei Pk + eD =− σ Ei Pk + .
2e ∂xi 2e e ∂xi
(22)

The spin-dependent potentials μs,k are defined by generalized Einstein equations


eD(∂(nPk )/∂μs,k ) = nμ. In relation with the continuity equation

h̄ ∂(nPk ) ∂js,ik h̄ (nPk )


+ =− (23)
2 ∂t ∂xi 2 τs

spin drift and diffusion can be predicted. τs denotes the spin relaxation time.
New phenomena occur, when spin-orbit coupling is taken into account [182,184,
185]. It was already shown in [182] that the spin-orbit interaction couples the charge
and spin current densities in such a way that a term proportional to −(2e/h̄)ikl js,kl
is added to Eq. (20) and a term proportional to (h̄/(2e))ikl jl is added to Eq. (22).
The different signs follow from the fact that the charge current density changes sign
under both space and time inversion, whereas the spin current density changes sign
only under space inversion. This yields the equations [183, 185]:
 
j = σ E + eD∇n + αSH σ (E × P ) + αSH eD ∇ × (nP ) (24)
  
h̄ ∂(nPk ) ∂n
js,ik = − σ Ei Pk + eD − αSH σ ikl El − αSH eDikl .
2e ∂xi ∂xl
(25)

ikl denotes the totally antisymmetric tensor; αSH is the spin Hall angle defined
as the ratio of the spin Hall conductivity σs to the diagonal charge conductivity
σ : αSH = (2h̄/e)(σs /σ ); for a tabulation of αSH values, see [185].
The cross terms appearing in Eqs. (24), (25) relate to various Hall effect
phenomena. The third term on the r.h.s. of Eq. (24) yields a charge current density
proportional to both applied electric field and spin polarization, i.e., in the presence
of a finite magnetization, an applied electric field induces a Hall effect proportional
to the magnetization component perpendicular to the applied electric field. This is
exactly the anomalous Hall effect discussed in section “Hall Effect ” and described
phenomenologically by Eq. (19).

Spin Hall Effect

The third and fourth term on the r.h.s. of Eq. (25) lead to the spin Hall effect
[184], in which a charge current produces a perpendicular spin current that leads
to spin accumulation at the boundaries of a finite sample. Consider a narrow flat
metal or semiconductor strip of width w (−∞ < x < ∞, −w/2 ≤ y ≤ w/2,
0 ≤ z ≤ d) in an electric field E applied along the x-direction. Solving Eqs. (23),
(25) with the boundary condition of vanishing spin current density at the surface
462 M. Ziese

yields a√spin polarization density nPz = (αSH σ λs )/(eD)E sinh(y/λs ), where


λs = Dτs denotes the spin diffusion length. At the two sample edges, spin
polarization of opposite direction perpendicular to the sample is created within a
layer of thickness λs . The spin Hall effect was detected optically [186] in GaAs
and electrically in Al [188]. Figure 17a shows results obtained on an n-doped GaAs
film: At the sample edges, Kerr rotation signals with an extent of about 10 μm and
opposite polarity appear in zero magnetic field. A transverse magnetic field destroys
the spin polarization, i.e., the spin Hall effect – contrary to the normal Hall effect –
is a zero magnetic field phenomenon.

Inverse Spin Hall Effect

The fourth term on the r.h.s. of Eq. (24) leads to the inverse spin Hall effect, in
which a spin current produces a Hall voltage by the accumulation of charges at
the boundaries of a finite sample. Consider a narrow metallic or semiconducting
half-slab (−∞ < x < ∞, −w/2 ≤ y ≤ w/2, z ≤ 0). At the top surface, spins
are injected with a spin polarization along the x-axis. This leads to a nonzero spin
current density component js,zx ; solving Eqs. (25), (23) one finds a spin polarization
decreasing exponentially into the material. Since there is no charge flow through the
surfaces, the charge current generated by ∇ × P must be balanced by an induced
electric field along the y-direction: Ey = −αSH (eD)/(σ λs )(nPx0 ) exp(z/λs ); Px0
denotes the spin polarization at the top surface. The inverse spin Hall effect was
detected in Pt when spin pumping from an adjacent permalloy layer [189] or by
spin diffusion from a Cu-permalloy injector [187]; however, it does not only occur
in non-magnetic metals with strong spin-orbit interaction, but also in ferromagnetic
permalloy, after spin injection from an adjacent yttrium iron garnet (YIG) [190].
Figure 17d shows the experiment after [187]: spins are injected from a permalloy
pad into a Cu junction by driving a current through the ferromagnetic pad along one
of the Cu arms. The spin density diffuses further along the Cu arm into the Pt wire,
where it creates a Hall voltage by the inverse spin Hall effect.

Quantum Spin Hall Effect

The classical spin current phenomenology was extended by the discovery of the
quantum spin Hall effect in two-dimensional systems subjected to a strong magnetic
field and the realization that the spin Hall conductance σs = (h̄/(2e))αSH σ may be
quantized in units of e/(2π ) [191]. In this case, spin-up and spin-down electrons
move with opposite chirality in one-dimensional edge states around the sample,
e.g., in HgTe quantum wells [192]. A generalization of these observations leads
to the concept of the topological insulator with topologically protected conducting
surface states around insulating bulk states [193].
9 Magnetotransport

Fig. 17 Spin Hall effect [186]: (a) GaAs film with electrodes; a local Kerr effect probe measures the Kerr rotation across the sample; the Kerr rotation is
proportional to the spin polarization. (b) Kerr rotation as a function of position across the sample and magnetic field applied parallel to y. The concentration
at the sample edges and the dephasing due to the Hanle effect are clearly seen. (c) Kerr rotation as a function of position across the film. The solid line is a fit
to the sinh function derived in the text. Inverse spin Hall effect [187]: (d) Pt film connected via a Cu junction to a permalloy (Py) spin injector. A current I is
driven through permalloy and junction by the voltage Vs ; the inverse spin Hall voltage VH across the Pt strip is measured. (e) and (f) inverse spin Hall resistance
VH /I at 77 K and 300 K as a function of applied magnetic field
463
464 M. Ziese

Magnetoimpedance

Magnetoimpedance describes the dependence of the impedance Z of a soft ferro-


magnetic wire or ribbon on a static magnetic field. In soft ferromagnets such as
FeCoSiB amorphous wires, the effect can be substantial for frequencies in the high
kHz and low MHz regime and was therefore called giant magnetoimpedance (GMI)
[194, 195]. The basis for GMI is the classical skin effect occurring in conductors
at high frequencies. Let us consider a soft ferromagnetic ribbon with permeability
μ = μr μ0 such that the magnetic induction B is given in terms of the magnetic field
by B = μH . If the ribbon is conducting, and Maxwell’s equations can be solved
for a current Iac = I0 exp(iωt) flowing along the long ribbon axis (infinite slab of
thickness d, −d/2 ≤ x ≤ d/2):

(1 + i)u R
Z=R = . (26)
tanh [(1 + i)u] μac

√ the dc resistance of the ribbon and u = d/(2δ) with the skin depth
R denotes
δ = 2/(ωσ μ0 μr ). The second equation on the r.h.s. is valid for a material
with isotropic relative permeability μr . μac denotes the ac-susceptibility of the
ribbon, caused by the skin effect, when placed in an alternating magnetic field
Hac = H0 exp(iωt) along the ribbon axis. For small skin depths, d/δ  1, μac ∼
(δ/d)(1 − i), and the impedance is much larger than the dc resistance. When a static
magnetic field is applied, the relative permeability μr is substantially decreased
and the skin depth increased, thus leading to a large decrease of the impedance,
i.e., in the zeroth approximation, giant magnetoimpedance can be considered as a
magnetically controlled skin effect.
Beyond this basic approximation, various aspects have to be taken into account
[196]: in eddy current models, the role of different domain structures (circular
domains, stripe domains) on the microscopic eddy current loss was studied [197];
in domain models, the dispersion of the circumferential permeability [198] and the
different contributions from domain wall motion and magnetization rotation [199]
were clarified. Eddy current and domain models successfully explain the relevant
features of giant magnetoimpedance up to a frequency of about 100 MHz. In the
GHz range, the magnetoimpedance is dominated by ferromagnetic resonance effects
[200].

Measurements

The measurement considerations in this section are restricted to homogeneous films


of thickness d; for anisotropic films, see [201]. In two-contact measurements, the
resistance of the wires, the contacts, and the sample are measured in series; this is
9 Magnetotransport 465

Fig. 18 (a) Four-contact configuration for measurement of the longitudinal voltage V and the Hall
voltage VH . (b) van der Pauw geometries. (c) Corbino disk

acceptable, when the sample resistance dominates, e.g., with metallic samples and
soldered or welded contacts. CPP measurements on metallic multilayers are made
in two-contact configuration using superconducting contacts. Often, however, the
contact resistance is high or is even non-ohmic, when pn or Schottky junctions
are formed between contact material and sample. In these cases, two-contact
measurements should be avoided; care should always be taken in choosing the
appropriate contact material.
Consider first the four-contact configuration shown in Fig. 18a. When a current I
is injected through a tip into the sample, the current density j flows radially outward
from the tip and at distance r is given by

I 1 1
j = 2
r = E = − ∇Φ (27)
2π r d ρ ρ

where r is a radius vector within the plane of the sample and Φ denotes the electric
potential. The latter is found by integration; in the situation shown in Fig. 18a with
current I injected at 1 and extracted at 4, the potential difference between 2 and 3 is
given by
 
ρI r2
V = ln . (28)
πd r1

From this, the resistance R = V /I and the resistivity ρ might be obtained. The Hall
voltage is given by VH = ρyx I /d, where ρyx denotes the off-diagonal resistivity;
see Eq. (7). In this geometry, care has to be taken to ensure a careful alignment of the
Hall voltage contacts; any horizontal shift will lead to the addition of a longitudinal
voltage contribution to the Hall voltage.
In order to minimize the influence of contact size and placement, van der Pauw
suggested an alternative method for resistance measurements of thin homogeneous
samples with arbitrary shape [202]. Although this method is valid for arbitrarily
formed samples, the recommended shapes are shown in Fig. 18b, since these
466 M. Ziese

minimize the influence of the size and exact location of the contacts. Consider a
sample with four contacts 1, 2, 3, 4 arranged counterclockwise around its rim, see
Fig. 18b, Ikl be a current injected into contact k and extracted from contact l; in
case of dc currents eight, in case of ac currents four non-diagonal combinations
Ikl , k, l = 1, 2, 3, 4, can be defined. The voltages measured between contacts i
and j are labeled Vij . Then eight (four) voltage measurements yield eight (four)
positive resistance values: Rij,kl = Vij /Ikl . Measurement consistency requires that
Rij,kl = Rj i,lk . The reciprocity theorem [203] requires that R21,34 + R12,43 =
R43,12 + R34,21 , R32,41 + R23,14 = R14,23 + R41,32 . The sheet resistance can be
calculated from the two averaged resistances (dc case) RA = (R21,34 + R12,43 +
R43,12 + R34,21 )/4 and RB = (R32,41 + R23,14 + R14,23 + R41,32 )/4 using the van
der Pauw relation
   
πd πd
exp − RA + exp − RB = 1 . (29)
ρ ρ

If the sample has a line of symmetry, RA = RB , and

πd
ρ= RA . (30)
ln(2)

In case of Hall measurements, the procedure is similar; with use of ac currents, the
diagonal resistances R24,13 and R13,24 are measured. The Hall resistivity loop as a
function of applied magnetic field is then obtained as

ρyx = (R24,13 − R13,24 )d . (31)

If one is only interested in the Hall constant, resistivity values for positive and
negative magnetic field should be averaged.
Hall effect measurements are only possible, when the sample has edges placed
in such a way that charges accumulate. This was the case in the configurations
discussed so far. These correspond to a vanishing current density component
perpendicular to the longitudinal current density component. Measurements in the
absence of a transverse electric field, i.e., a Hall field, can be performed in the
so-called Corbino disk configuration; see Fig. 18c. Here the conducting annulus
has contacts around the outer and inner rims. In the absence of a magnetic field,
the current density flows in radial direction; in a perpendicular magnetic field, the
current density flows in spirals. In this configuration, the conductance related to σxx
is measured.
In case of measurements of the spin accumulation, nonlocal resistivity measure-
ments are performed [55]. Usually four contacts are defined on the sample in order
to perform local four-point measurements. Further contacts are defined outside the
current path for the measurement of voltages generated by the diffusion of carriers.
Care has to be taken to define a unique ground potential in these measurements
[204].
9 Magnetotransport 467

References
1. Ohm, G.S.: Die galvanische Kette. T. H. Riemann, Berlin (1827)
2. Onsager, L.: Reciprocal relations in irreversible processes. Phys. Rev. 37, 405 (1931)
3. Lide, D.R. (ed.): CRC Handbook of Chemistry and Physics, 84th edn. CRC Press, Boca
Raton (2003)
4. Ho, C.Y., Ackerman, M.W., Wu, K.Y., Havill, T.N., Bogaard, R.H., Matula, R.A., Oh, S.G.,
James, H.M.: Electrical resistivity of ten selected binary alloy systems. J. Phys. Chem. Ref.
Data 12, 183 (1983)
5. Ashcroft, N.W., Mermin, N.D.: Solid State Physics. Holt-Saunders Japan, Tokyo (1981)
6. Drude, P.: Zur Elektronentheorie der Metalle. Ann. Phys. 306, 566 (1900)
7. Matthiessen, A., von Bose, M.: On the influence of temperature on the electric conducting
power of metals. Phil. Trans. R. Soc. Lond. 152, 1 (1862)
8. Gerlach, W., Schneiderhan, K.: Ferromagnetismus und elektrische Eigenschaften. Wider-
stand, magnetische Widerstandsänderung und wahre Magnetisierung beim Curie-Punkt. Ann.
Phys. 398, 772 (1930)
9. Mott, N.F.: Electrons in disordered structures. Adv. Phys. 16, 49 (1967)
10. Anderson, P.W.: Absence of diffusion in certain random lattices. Phys. Rev. 109, 1492 (1958)
11. Mooij, J.H.: Electrical conduction in concentrated disordered transition metal alloys. Phys.
Stat. Sol. (A) 17, 521 (1973)
12. Fisk, Z., Webb, G.W.: Saturation of the high-temperature normal-state electrical resistivity of
superconductors. Phys. Rev. Lett. 36, 1084 (1976)
13. Millis, A.J., Hu, J., Das Sarma, S.: Resistivity saturation revisited: results from a dynamical
mean field theory. Phys. Rev. Lett. 82, 2354 (1999)
14. Lee, P.A., Ramakrishnan, T.V.: Disordered electronic systems. Rev. Mod. Phys. 57, 287
(1985)
15. Bergmann, G.: Consistent temperature and field dependence in weak localization. Phys. Rev.
B 28, 515 (1983)
16. Campbell, I.A., Fert, A.: Transport properties of ferromagnets. In: Wohlfarth, E.P.
(ed.) Ferromagnetic Materials, vol. 3, pp. 751–800. North-Holland Publishing Company,
Amsterdam (1982)
17. Raquet, B., Viret, M., Sondergard, E., Céspedes, O., Mamy, R.: Electron-magnon scattering
and magnetic resistivity in 3d ferromagnets. Phys. Rev. B 66, 024433 (2002)
18. Du, X., Tsai, S.-W., Maslov, D.L., Hebard, A.F.: Metal-insulator-like behavior in semimetallic
bismuth and graphite. Phys. Rev. Lett. 94, 166601 (2005)
19. Kempa, H., Kopelevich, Y., Mrowka, F., Setzer, A., Torres, J.H.S, Höhne, R., Esquinazi,
P.: Magnetic-field-driven superconductor-insulator-type transition in graphite. Solid State
Commun. 115, 539 (2000)
20. Kohler, M.: Zur magnetischen Widerstandsänderung reiner Metalle. Ann. Phys. 424, 211
(1938)
21. Beer, A.C.: Galvanomagnetic Effects in Semiconductors. Academic Press, New York (1963)
22. Xu, R., Husmann, A., Rosenbaum, T.F., Saboungi, M.-L., Enderbya, J.E., Littlewood, P.B.:
Large magnetoresistance in non-magnetic silver chalcogenides. Nature 390, 57 (1997)
23. McClure, J.W., Spry, W.J.: Linear magnetoresistance in the quantum limit in graphite. Phys.
Rev. 165, 809 (1968)
24. Friedman, A.L., Tedesco, J.L., Campbell, P.M., Culbertson, J.C., Aifer, E., Perkins, F.K.,
Myers-Ward, R.L., Hite, J.K., Eddy, C.R. Jr., Jernigan, G.G., Gaskill, D.K.: Quantum linear
magnetoresistance in multilayer epitaxial graphene. Nano Lett. 10, 3962 (2010)
25. Abrikosov, A.A.: Quantum magnetoresistance. Phys. Rev. B 58, 2788 (1998)
26. Argyres, P.N.: Quantum theory of longitudinal magneto-resistance. J. Phys. Chem. Solids 4,
19 (1958)
27. Adams, E.N., Holstein, T.D.: Quantum theory of transverse galvano-magnetic phenomena. J.
Phys. Chem. Solids 10, 254 (1959)
468 M. Ziese

28. Thomson, W.: On the electro-dynamic qualities of metals: effects of magnetization on the
electric conductivity of nickel and of iron. Proc. R. Soc. (Lond.) 8, 546 (1857)
29. McGuire, T.R., Potter, R.I.: Anisotropic magnetoresistance in ferromagnetic 3d alloys. IEEE
Trans. Mag. 11, 1018 (1975)
30. Döring, W.: Die Abhängigkeit des Widerstandes von Nickelkristallen von der Richtung der
spontanen Magnetisierung. Ann. Phys. 424, 259 (1938)
31. Döring, W., Simon, G.: Die Richtungsabhängigkeit der Magnetostriktion. Ann. Phys. 460,
373 (1960)
32. Birss, R.R.: Symmetry and Magnetism. North-Holland, Amsterdam (1964)
33. Ziese, M., Vrejoiu, I., Hesse, D.: Structural symmetry and magnetocrystalline anisotropy of
SrRuO3 films on SrTiO3 . Phys. Rev. B 81, 184418 (2010)
34. Tang, H.X., Kawakami, R.K., Awschalom, D.D., Roukes, M.L.: Giant planar Hall effect in
epitaxial (Ga,Mn)As devices. Phys. Rev. Lett. 90, 107201 (2003)
35. Bason, Y., Klein, L., Yau, J.-B., Hong, X., Ahn, C.H.: Giant planar Hall effect in colossal
magnetoresistive La0.84 Sr0.16 MnO3 thin films. Appl. Phys. Lett. 84, 2593 (2004)
36. Mott, N.F.: The electrical conductivity of transition metals. Proc. R. Soc. 153, 699 (1936)
37. Mott, N.F.: The resistance and thermoelectric properties of the transition metals. Proc. R.
Soc. 156, 368 (1936)
38. Dorleijn, J.W.F., Miedema, A.R.: A quantitative investigation of the two current conduction
in nickel alloys. J. Phys. F 5, 487 (1975)
39. Dorleijn, J.W.F.: Electrical conduction in ferromagnetic metals. Philips Res. Rep. 31, 287
(1976)
40. Fert, A., Campbell, I.A.: Electrical resistivity of ferromagnetic nickel and iron based alloys.
J. Phys. F 6, 849 (1976)
41. Ross, R.N., Price, D.C., Williams, G.: Resistivity and magnetoresistance anisotropy
associated with s, p impurities in ferromagnetic Ni and Co. J. Phys. F 8, 2367 (1978)
42. Ross, R.N., Price, D.C., Williams, G.: Spontaneous magnetoresistance anisotropy and
deviations from Matthiessen’s rule induced by 5s, p impurities in Fe. J. Magn. Magn. Mater.
10, 59 (1979)
43. Smit, J.: Magnetoresistance of ferromagnetic metals and alloys at low temperatures. Physica
17, 612 (1951)
44. Campbell, I.A., Fert, A., Jaoul, O.: The spontaneous resistivity anisotropy in Ni-based alloys.
J. Phys. C 3, S95 (1970)
45. Malozemoff, A.P.: Anisotropic magnetoresistance with cubic anisotropy and weak ferromag-
netism: a new paradigm. Phys. Rev. B 34, 1853 (1986)
46. Jaoul, O., Campbell, I.A., Fert, A.: Spontaneous resistivity anisotropy in Ni alloys. J. Magn.
Magn. Mater. 5, 23 (1977)
47. Grünberg, P., Schreiber, R., Pang, Y., Brodsky, M.B., Sowers, H.: Layered magnetic
structures: evidence for antiferromagnetic coupling of Fe layers across Cr interlayers. Phys.
Rev. Lett. 57, 2442 (1986)
48. Baibich, M.N., Broto, J.M., Fert, A., Van Dau, F.N., Petroff, F., Eitenne, P., Creuzet,
G., Friederich, A., Chazelas, J.: Giant magnetoresistance of (001)Fe/(001)Cr magnetic
superlattices. Phys. Rev. Lett. 61, 2472 (1988)
49. Binasch, G., Grünberg, P., Saurenbach, F., Zinn, W.: Enhanced magnetoresistance in layered
magnetic structures with antiferromagnetic interlayer exchange. Phys. Rev. B 39, 4828 (1989)
50. Dieny, B.: Giant magnetoresistance in spin-valve multilayers. J. Magn. Magn. Mater. 136,
335 (1994)
51. Parkin, S.S.P., More, N., Roche, K.P.: Oscillations in exchange coupling and magnetoresis-
tance in metallic superlattice structures: Co/Ru, Co/Cr, and Fe/Cr. Phys. Rev. Lett. 64, 2304
(1990)
52. Edwards, D.M., Mathon, J., Muniz, R.B.: A resistor network theory of the giant magnetore-
sistance in magnetic superlattices. IEEE Trans. Mag. 21, 3548 (1991)
53. Mathon, J.: Exchange interactions and giant magnetoresistance in magnetic multilayers.
Contemp. Phys. 32, 143 (1991)
9 Magnetotransport 469

54. Mathon, J.: Phenomenological theory of giant magnetoresistance. In: Ziese, M., Thornton,
M.J. (eds.) Spin Electronics, pp. 71–88. Springer, Berlin/Heidelberg (2001)
55. Johnson, M., Silsbee, R.H.: Spin-injection experiment. Phys. Rev. B 37, 5326 (1988)
56. van Son, P.C., van Kempen, H., Wyder, P.: Boundary resistance of the ferromagnetic-
nonferromagnetic metal interface. Phys. Rev. Lett. 58, 2271 (1987)
57. Valet, T., Fert, A.: Theory of the perpendicular magnetoresistance in magnetic multilayers.
Phys. Rev. B 48, 7099 (1993)
58. Park, B.G., Wunderlich, J., Martí, X., Holý, V., Kurosaki, Y., Yamada, M., Yamamoto, H.,
Nishide, A., Hayakawa, J., Takahashi, H., Shick, A.B., Jungwirth, T.: A spin-valve-like
magnetoresistance of an antiferromagnet-based tunnel junction. Nat. Mater. 10, 347 (2011)
59. Gould, C., Rüster, C., Jungwirth, T., Girgis, E., Schott, G.M., Giraud, R., Brunner, K.,
Schmidt, G., Molenkamp, L.W.: Tunneling anisotropic magnetoresistance: a spin-valve-like
tunnel magnetoresistance using a single magnetic layer. Phys. Rev. Lett. 93, 117203 (2004)
60. von Helmolt, R., Wecker, J., Holzapfel, B., Schultz, L., Samwer, K.: Giant negative
magnetoresistance in perovskitelike La2/3 Ba1/3 MnOx ferromagnetic films. Phys. Rev. Lett.
71, 2331 (1993)
61. Jin, S., Tiefel, T.H., McCormack, M., Fastnacht, R.A., Ramesh, R., Chen, L.H.: Thousandfold
change in resistivity in magnetoresistive La-Ca-Mn-O films. Science 264, 413 (1994)
62. Searle, C.W., Wang, S.T.: Studies of the ionic ferromagnet (LaPb)MnO3 . Electric transport
and ferromagnetic properties. Can. J. Phys. 48, 2023 (1970)
63. Fontcuberta, J., Martínez, B., Seffar, A., Piñol, S., García-Muñoz, J.L., Obradors, X.: Colossal
magnetoresistance of ferromagnetic manganites: structural tuning and mechanisms. Phys.
Rev. Lett. 76, 1122 (1996)
64. Ziese, M., Srinitiwarawong, C.: Extraordinary Hall effect in La0.7 Ca0.3 MnO3 and
La0.7 Ba0.3 MnO3 thin films. Europhys. Lett. 45, 256 (1999)
65. Ramirez, A.P.: Colossal magnetoresistance. J. Phys.: Condens. Matter 9, 8171 (1997)
66. Fujishiro, H., Fukase, T., Ikebe, M.: Charge ordering and sound velocity anomaly in
La1−x Srx MnO3 (x ≥ 0.5). J. Phys. Soc. Jpn. 67, 2582 (1998)
67. Shenoy, V.B., Rao, C.N.R.: Electronic phase separation and other novel phenomena and
properties exhibited by mixed-valent rare-earth manganites and related materials. Phil. Trans.
R. Soc. A 366, 63 (2008)
68. Jonker, G., van Santen, J.: Ferromagnetic compounds of manganese with perovskite structure.
Physica 16, 337 (1950)
69. Coey, J.M.D., Viret, M., von Molnár, S.: Mixed-valence manganites. Adv. Phys. 48, 167
(1999)
70. Zener, C.: Interaction between the d-shells in the transition metals. Ferromagnetic compounds
of manganese with perovskite structure. Phys. Rev. 82, 403 (1951)
71. de Gennes, P.G.: Effects of double exchange in magnetic crystals. Phys. Rev. 118, 141 (1960)
72. Böttcher, D., Henk, J.: Magnetic properties of strained La2/3 Sr1/3 MnO3 perovskites from
first principles. J. Phys.: Condens. Matter 25, 136005 (2013)
73. Salomon, M.B., Jaime, M.: The physics of manganites: structure and transport. Rev. Mod.
Phys. 73, 583 (2001)
74. Millis, A.J., Littlewood, P.B., Shraiman, B.I.: Double exchange alone does not explain the
resistivity of La1−x Srx MnO3 . Phys. Rev. Lett. 74, 5144 (1995)
75. J. M. D Coey, Viret, M., Ranno, L., and Ounadjela, K.: Electron localization in mixed-valence
manganites. Phys. Rev. Lett. 75, 3910 (1995)
76. Dagotto, E., Hotta, T., Moreo, A.: Colossal magnetoresistant materials: the key role of phase
separation. Phys. Rep. 344, 1 (2001)
77. Ziese, M. and Vrejoiu, I.: Properties of manganite/ruthenate superlattices with ultrathin
layers. Phys. Status Solidi RRL 7, 243 (2013)
78. de Groot, R.A., Mueller, F.M., van Engen, P.G., Buschow, K.H.J.: New class of materials:
half-metallic ferromagnets. Phys. Rev. Lett. 50, 2024 (1983)
79. Hwang, H.Y., Cheong, S.-W., Ong, N.P., Batlogg, B.: Spin-polarized intergrain tunneling in
La2/3 Sr1/3 MnO3 . Phys. Rev. Lett. 77, 2041 (1996)
470 M. Ziese

80. Gupta, A., Gong, G.Q., Xiao, G., Duncombe, P.R., Lecoeur, P., Trouilloud, P., Wang, Y.Y.,
Dravid, V.P., Sun, J.Z.: Grain-boundary effects on the magnetoresistance properties of
perovskite manganite films. Phys. Rev. B 54, R15629 (1996)
81. Ziese, M.: Extrinsic magnetotransport phenomena in ferromagnetic oxides. Rep. Prog. Phys.
65, 143 (2002)
82. Tedrow, P.M., Meservey, R.: Spin-dependent tunneling into ferromagnetic nickel. Phys. Rev.
Lett. 26, 192 (1971)
83. Tedrow, P.M., Meservey, R.: Spin polarization of electrons tunneling from films of Fe, Co,
Ni, and Gd. Phys. Rev. B 7, 318 (1973)
84. Meservey, R., Tedrow, P.M.: Spin-polarized electron tunneling. Phys. Rep. 238, 173 (1994)
85. Soulen, R.J. Jr., Byers, J.M., Osofsky, M.S., Nadgorny, B., Ambrose, T., Cheng, S.F.,
Broussard, P.R., Tanaka, C.T., Nowak, J., Moodera, J.S., Barry, A., Coey, J.M.D.: Measuring
the spin polarization of a metal with a superconducting point contact. Science 282, 85 (1998)
86. Bugoslavsky, Y., Miyoshi, Y., Clowes, S.K., Branford, W.R., Lake, M., Brown, I., Caplin,
A.D., Cohen, L.F.: Possibilities and limitations of point-contact spectroscopy for measure-
ments of spin polarization. Phys. Rev. B 71, 104523 (2005)
87. Mazin, I.I.: How to define and calculate the degree of spin polarization in ferromagnets. Phys.
Rev. Lett. 83, 1427 (1999)
88. Julliere, M.: Tunneling between ferromagnetic films. Phys. Lett. 54A, 225 (1975)
89. Moodera, J.S., Kinder, L.R., Wong, T.M., Meservey, R.: Large magnetoresistance at room
temperature in ferromagnetic thin film tunnel junctions. Phys. Rev. Lett. 74, 3273 (1995)
90. Åkerman, J.J., Slaughter, J.M., Dave, R.W., Schuller, I.K.: Tunneling criteria for magnetic-
insulator-magnetic structures. Appl. Phys. Lett. 79, 3104 (2001)
91. Bowen, M., Bibes, M., Barthélémy, A., Contour, J.-P., Anane, A., Lemaître, Y., Fert, A.:
Nearly total spin polarization in La2/3 Sr1/3 MnO3 from tunneling experiments. Appl. Phys.
Lett. 82, 233 (2003)
92. Faure-Vincent, J., Tiusan, C., Jouguelet, E., Canet, F., Sajieddine, M., Bellouard, C., Popova,
E., Hehn, M., Montaigne, F., Schuhl, A.: High tunnel magnetoresistance in epitaxial
Fe/MgO/Fe tunnel junctions. Appl. Phys. Lett. 82, 4507 (2003)
93. Johnson, P.D., Güntherodt, G.: Spin-polarized photoelectron spectroscopy as a probe of
magnetic systems. In: Kronmüller, H., Parkin, S.S.P. (eds.) Handbook of Magnetism and
Advanced Magnetic Materials, pp. 1635–1657. Wiley, Chichester (2007)
94. Monsma, D.J., Parkin, S.S.P.: Spin polarization of tunneling current from ferromagnet/Al2 O3
interfaces using copper-doped aluminum superconducting films. Appl. Phys. Lett. 77, 720
(2000)
95. Yates, K.A., Branford, W.R., Magnus, F., Miyoshi, Y., Morris, B., Cohen, L.F., Sousa, P.M.,
Conde, O., Silvestre, A.J.: The spin polarization of CrO2 revisited. Appl. Phys. Lett. 91,
172504 (2007)
96. Hu, G., Suzuki, Y.: Negative spin polarization of Fe3 O4 in magnetite/manganite-based
junctions. Phys. Rev. Lett. 89, 276601 (2002)
97. Nadgorny, B., Osofsky, M.S., Singh, D.J., Woods, G.T., Soulen, R.J. Jr., Lee, M.K., Bu, S.D.,
Eom, C.B.: Measurements of spin polarization of epitaxial SrRuO3 thin films. Appl. Phys.
Lett. 82, 427 (2003)
98. Sinković, B., Shekel, E., Hulbert, S.L.: Spin-resolved iron surface density of states. Phys.
Rev. B 52, R8696 (1995)
99. Rampe, A., Hartmann, D., Weber, W., Popovic, S., Reese, M., Güntherodt, G.: Induced spin
polarization and interlayer exchange coupling of the systems Rh/Co(0001) and Ru/Co(0001).
Phys. Rev. B 51, 3230 (1995)
100. Eib, W., Alvarado, S.F.: Spin-polarized photoelectrons from nickel single crystals. Phys. Rev.
Lett. 37, 444 (1976)
101. Park, J.-H., Vescovo, E., Kim, H.-J., Kwon, C., Ramesh, R., Venkatesan, T.: Magnetic
properties at surface boundary of a half-metallic ferromagnet La0.7 Sr0.3 MnO3 . Phys. Rev.
Lett. 81, 1953 (1998)
9 Magnetotransport 471

102. Dedkov, Y.S., Fonin, M., König, C., Rüdiger, U., Güntherodt, G., Senz, S., Hesse, D.: Room-
temperature observation of high-spin polarization of epitaxial CrO2 (100) island films at the
fermi energy. Appl. Phys. Lett. 80, 4181 (2002)
103. Fonin, M., Pentcheva, R., Dedkov, Y.S., Sperlich, M., Vyalikh, D.V., Scheffler, M., Rüdiger,
U., Güntherodt, G.: Surface electronic structure of the Fe3 O4 (100): evidence of a half-metal
to metal transition. Phys. Rev. B 72, 104436 (2005)
104. Tsymbal, E.Y., Mryasov, O.N., LeClair, P.R.: Spin-dependent tunnelling in magnetic tunnel
junction. J. Phys.: Condens. Matter 15, R109 (2003)
105. Moodera, J.S., Nowak, J., van de Veerdonk, R.J.M.: Interface magnetism and spin wave
scattering in ferromagnet-insulator-ferromagnet tunnel junctions. Phys. Rev. Lett. 80, 2941
(1998)
106. Brinkman, W.F., Dynes, R.C., Rowell, J.M.: Tunneling conductance of asymmetrical barriers.
J. Appl. Phys. 41, 1915 (1970)
107. De Teresa, J.M., Barthélémy, A., Fert, A., Contour, J.P., Lyonnet, R., Montaigne, F., Seneor,
P., Vaurès, A.: Inverse tunnel magnetoresistance in Co/SrTiO3 /La0.7 Sr0.3 MnO3 : new ideas
on spin-polarized tunneling. Phys. Rev. Lett. 82, 4288 (1999)
108. Jo, M.-H., Mathur, N.D., Todd, N.K., Blamire, M.G.: Very large magnetoresistance and
coherent switching in half-metallic manganite tunnel junctions. Phys. Rev. B 61, R14905
(2000)
109. Mathon, J., Umerski, A.: Theory of tunneling magnetoresistance of an epitaxial
Fe/MgO/Fe(001) junction. Phys. Rev. B 63, 220403(R) (2001)
110. Yuasa, S., Nagahama, T., Fukushima, A., Suzuki, Y., Ando, K.: Giant room-temperature
magnetoresistance in single-crystal Fe/MgO/Fe magnetic tunnel junctions. Nat. Mater. 3, 868
(2004)
111. Parkin, S.P.P., Kaiser, C., Panchula, A., Rice, P.M., Hughes, B., Samant, M., Yang, S.-H.:
Giant tunnelling magnetoresistance at room temperature with MgO (100) tunnel barriers.
Nat. Mater. 3, 862 (2004)
112. Djayaprawira, D.D., Tsunekawa, K., Nagai, M., Maehara, H., Yamagata, S., Watanabe, N.,
Yuasa, S., Suzuki, Y., Ando, K.: 230% room-temperature magnetoresistance in CoFeB/M-
gO/CoFeB magnetic tunnel junctions. Appl. Phys. Lett. 86, 092502 (2005)
113. Kaiser, C., Panchula, A.F., Parkin, S.S.P.: Finite tunneling spin polarization at the compensa-
tion point of rare-earth-metal-transition-metal alloys. Phys. Rev. Lett. 95, 047202 (2005)
114. Wiesendanger, R.: Single-atom magnetometry. Curr. Opin. Solid State Mater. Sci. 15, 1
(2011)
115. Ding, H.F., Wulfhekel, W., Kirschner, J.: Ultra sharp domain walls in the closure domain
pattern of Co(0001). Europhys. Lett. 57, 100 (2002)
116. Oka, H., Tao, K., Wedekind, S., Rodary, G., Stepanyuk, V.S., Sander, D., Kirschner, J.:
Spatially modulated tunnel magnetoresistance on the nanoscale. Phys. Rev. Lett. 107, 187201
(2011)
117. Ferriani, P., von Bergmann, K., Vedmedenko, E.Y., Heinze, S., Bode, M., Heide, M.,
Bihlmayer, G., Blügel, S., Wiesendanger, R.: Atomic-scale spin spiral with a unique rotational
sense: Mn monolayer on W(001). Phys. Rev. Lett. 101, 027201 (2008)
118. Heinze, S., von Bergmann, K., Menzel, M., Brede, J., Kubetzka, A., Wiesendanger, R.,
Bihlmayer, G., Blügel, S.: Spontaneous atomic-scale magnetic skyrmion lattice in two
dimensions. Nat. Phys. 7, 713 (2011)
119. Coey, J.M.D., Berkowitz, A.E., Balcells, L., Putris, F.F., Parker, F.T.: Magnetoresistance of
magnetite. Appl. Phys. Lett. 72, 734 (1998)
120. Coey, J.M.D., Berkowitz, A.E., Balcells, L., Putris, F.F., Barry, A.: Magnetoresistance of
chromium dioxide powder compacts. Phys. Rev. Lett. 80, 3815 (1998)
121. Coey, J.M.D.: Powder magnetoresistance. J. Appl. Phys. 85, 5576 (1999)
122. Francis, T.L., Mermer, Ö., Veeraraghavan, G., Wohlgenannt, M.: Large magnetoresistance at
room temperature in semiconducting polymer sandwich devices. New J. Phys. 6, 185 (2004)
123. Bloom, F.L., Wagemans, W., Kemerink, M., Koopmans, B.: Separating positive and negative
magnetoresistance in organic semiconductor devices. Phys. Rev. Lett. 99, 257201 (2007)
472 M. Ziese

124. Bergeson, J.D., Prigodin, V.N., Lincoln, D.M., Epstein, A.J.: Inversion of magnetoresistance
in organic semiconductors. Phys. Rev. Lett. 100, 067201 (2008)
125. Hu, B., Wu, Y.: Tuning magnetoresistance between positive and negative values in organic
semiconductors. Nat. Mater. 6, 985 (2007)
126. Wohlgenannt, M., Bobbert, P.A., Koopmans, B.: Intrinsic magnetic field effects in organic
semiconductors. MRS Bull. 39, 590 (2014)
127. Majumdar, S., Lill, J.-O., Rajander, J., Majumdar, H.: Observation of ferromagnetic ordering
in conjugated polymers exhibiting OMAR effect. Org. Electron. 21, 66 (2015)
128. Landauer, R.: Conductance determined by transmission: probes and quantised constriction
resistance. J. Phys.: Condens. Matter 1, 8099 (1989)
129. Wharam, D.A., Thornton, T.J., Newbury, R., Pepper, M., Ahmed, H., Frost, J.E.F., Hasko,
D.G., Peacock, D.C., Ritchie, D.A., Jones, G.A.C.: One-dimensional transport and the
quantisation of the ballistic resistance. J. Phys. C: Solid State Phys. 21, L209 (1988)
130. Ohnishi, H., Kondo, Y., Takayanagi, K.: Quantized conductance through individual rows of
suspended gold atoms. Nature 395, 780 (1998)
131. Ono, T., Ooka, Y., Miyajima, H., Otani, Y.: 2e2 /h to e2 /h switching of quantum conductance
associated with a change in nanoscale ferromagnetic domain structure. Appl. Phys. Lett. 75,
1622 (1999)
132. García, N., Muñoz, M., Zhao, Y.-W.: Magnetoresistance in excess of 200% in ballistic ni
nanocontacts at room temperature and 100 oe. Phys. Rev. Lett. 82, 2923 (1999)
133. Versluijs, J.J., Bari, M.A., Coey, J.M.D.: Magnetoresistance of half-metallic oxide nanocon-
tacts. Phys. Rev. Lett. 87, 026601 (2001)
134. Chung, S.H., Muñoz, M., García, N., Egelhoff, W.F., Gomez, R.D.: Universal scaling of
ballistic magnetoresistance in magnetic nanocontacts. Phys. Rev. Lett. 89, 287203 (2002)
135. Céspedes, O., Watts, S.M., Coey, J.M.D., Dörr, K., Ziese, M.: Magnetoresistance and
electrical hysteresis in stable half-metallic La0.7 Sr0.3 MnO3 and Fe3 O4 nanoconstrictions.
Appl. Phys. Lett. 87, 083102 (2005)
136. Bruno, P.: Geometrically constrained magnetic wall. Phys. Rev. Lett. 83, 2425 (1999)
137. Tatara, G., Zhao, Y.-W., Muñoz, M., García, N.: Domain wall scattering explains 300%
ballistic magnetoconductance of nanocontacts. Phys. Rev. Lett. 83, 2030 (1999)
138. Gregg, J.F., Petej, I., Jouguelet, E., Dennis, C.: Spin electronics – a review. J. Phys. D: Appl.
Phys. 35, R121 (2002)
139. Yakushiji, K., Mitani, S., Ernult, F., Takanashi, K., Fujimori, H.: Spin-dependent tunneling
and coulomb blockade in ferromagnetic nanoparticles. Phys. Rep. 451, 1 (2007)
140. Yuasa, S., Nagahama, T., Suzuki, Y.: Spin-polarized resonant tunneling in magnetic tunnel
junctions. Science 297, 234 (2002)
141. Lüders, U., Bibes, M., Bouzehouane, K., Jacquet, E., Contour, J.-P., Fusil, S., Bobo, J.-F.,
Fontcuberta, J., Barthélémy, A., Fert, A.: Spin filtering through ferrimagnetic NiFe2 O4 tunnel
barriers. Appl. Phys. Lett. 88, 082505 (2006)
142. Garcia, V., Bibes, M., Bocher, L., Valencia, S., Kronast, F., Crassous, A., Moya, X., Enouz-
Vedrenne, S., Gloter, A., Imhoff, D., Deranlot, C., Mathur, N.D., Fusil, S., Bouzehouane, K.,
Barthélémy, A.: Ferroelectric control of spin polarization. Science 327, 1106 (2010)
143. Gajek, M., Bibes, M., Fusil, S., Bouzehouane, K., Fontcuberta, J., Barthélémy, A., Fert, A.:
Tunnel junctions with multiferroic barriers. Nat. Mater. 6, 296 (2007)
144. Schelp, L.F., Fert, A., Fettar, F., Holody, P., Lee, S.F., Maurice, J.L., Petroff, F., Vaurès, A.:
Spin-dependent tunneling with coulomb blockade. Phys. Rev. B 56, R5747 (1997)
145. Guéron, S., Deshmukh, M.M., Myers, E.B., Ralph, D.C.: Tunneling via individual electronic
states in ferromagnetic nanoparticle. Phys. Rev. Lett. 83, 4148 (1999)
146. Graf, H., Vancea, J., Hoffmann, H.: Single-electron tunneling at room temperature in cobalt
nanoparticles. Appl. Phys. Lett. 80, 1264 (2002)
147. Barnaś, J., Fert, A.: Magnetoresistance oscillations due to charging effects in double
ferromagnetic tunnel junctions. Phys. Rev. Lett. 80, 1058 (1998)
148. Barnaś, J., Fert, A.: Effects of spin accumulation on single-electron tunneling in a double
ferromagnetic microjunction. Europhys. Lett. 44, 85 (1998)
9 Magnetotransport 473

149. Barnaś, J., Fert, A.: Interplay of spin accumulation and coulomb blockade in double
ferromagnetic junctions. J. Magn. Magn. Mater. 192, L391 (1999)
150. Esaki, L.: New phenomenon in narrow Germanium p − n junctions. Phys. Rev. 109, 603
(1958)
151. Itoh, H., Inoue, J., Umerski, A., Mathon, J.: Quantum oscillation of magnetoresistance in
tunneling junctions with a nonmagnetic spacer. Phys. Rev. B 68, 174421 (2003)
152. Moodera, J.S., Santos, T.S., Nagahama, T.: The phenomena of spin-filter tunnelling. J. Phys.:
Condens. Matter 19, 165202 (2007)
153. Kok, M., Beukers, J.N., Brinkman, A.: Spin-polarized tunneling through a ferromagnetic
insulator. J. Appl. Phys. 105, 07C919 (2009)
154. Caffrey, N.M., Fritsch, D., Archer, T., Sanvito, S., Ederer, C.: Spin-filtering efficiency of
ferrimagnetic spinels CoFe2 O4 and NiFe2 O4 . Phys. Rev. B 87, 024419 (2013)
155. Santos, T.S., Moodera, J.S.: Observation of spin filtering with a ferromagnetic EuO tunnel
barrier. Phys. Rev. B 69, 241203(R) (2004)
156. Gajek, M., Bibes, M., Varela, M., Fontcuberta, J., Herranz, G., Fusil, S., Bouzehouane,
K., Barthélémy, A., Fert, A.: La2/3 Sr1/3 MnO3 -La0.1 Bi0.9 MnO3 heterostructures for spin
filtering. J. Appl. Phys. 99, 08E504 (2006)
157. Tsymbal, E.Y., Kohlstedt, H.: Tunneling across a ferroelectric. Science 313, 181 (2006)
158. Garcia, V., Bibes, M.: Ferroelectric tunnel junctions for information storage and processing.
Nat. Commun. 5, 4289 (2014)
159. Kohlstedt, H., Petraru, A., Szot, K., Rüdiger, A., Meuffels, P., Haselier, H., Waser, R.,
Nagarajan, V.: Method to distinguish ferroelectric from nonferroelectric origin in case of
resistive switching in ferroelectric capacitors. Appl. Phys. Lett. 92, 062907 (2008)
160. Pantel, D., Haidong, H., Goetze, S., Werner, P., Kim, D.J., Gruverman, A., Hesse, D.,
Alexe, M.: Tunnel electroresistance in junctions with ultrathin ferroelectric Pb(Zr0.2 Ti0.8 )O3
barriers. Appl. Phys. Lett. 100, 232902 (2012)
161. Pantel, D., Goetze, S., Hesse, D., Alexe, M.: Reversible electrical switching of spin
polarization in multiferroic tunnel junctions. Nat. Mater. 11, 289 (2012)
162. Quindeau, A., Fina, I., Marti, X., Apachitei, G., Ferrer, P., Nicklin, C., Pippel, E., Hesse, D.,
Alexe, M.: Four-state ferroelectric spin-valve. Sci. Rep. 5, 9749 (2015)
163. Béa, H., Gajek, M., Bibes, M., Barthélémy, A.: Spintronics with multiferroics. J. Phys.:
Condens. Matter 20, 434221 (2008)
164. Bibes, M., Villegas, J.E., Barthélémy, A.: Ultrathin oxide films and interfaces for electronics
and spintronics. Adv. Phys. 60, 5 (2011)
165. Fusil, S., Garcia, V., Barthélémy, A., Bibes, M.: Magnetoelectric devices for spintronics.
Annu. Rev. Mater. Res. 44, 91 (2014)
166. Hall, E.H.: On a new action of the magnet on electric currents. Am. J. Math. 2, 287 (1879)
167. Pugh, E.M., Rostoker, N.: Hall effect in ferromagnetic materials. Rev. Mod. Phys. 25, 151
(1953)
168. Ziese, M., Vrejoiu, I.: Anomalous and planar Hall effect of orthorhombic and tetragonal
SrRuO3 layers. Phys. Rev. B 84, 104413 (2011)
169. Matsuno, J., Ogawa, N., Yasuda, K., Kagawa, F., Koshibae, W., Nagaosa, N., Tokura, Y.,
Kawasaki, M.: Interface-driven topological Hall effect in SrRuO3 -SrIrO3 bilayer. Sci. Adv.
2, e1600304 (2016)
170. Hurd, C.M.: Pressing electricity. In: Chien, C.L., Westgate, C.R. (eds.) The Hall Effect and
Its Applications, pp. 1–54. Plenum Press, New York (1980)
171. Smit, J.: The spontaneous Hall effect in ferromagnetics I. Physica 21, 877 (1955)
172. Smit, J.: The spontaneous Hall effect in ferromagnetics II. Physica 24, 39 (1958)
173. Berger, L.: Side-jump mechanism for the Hall effect of ferromagnets. Phys. Rev. B 2, 4559
(1970)
174. Lyanda-Geller, Y., Chun, S.H., Salamon, M.B., Goldbart, P.M., Han, P.D., Tomioka, Y.,
Asamitsu, A., Tokura, Y.: Charge transport in manganites: hopping conduction, the anomalous
hall effect, and universal scaling. Phys. Rev. B 63, 184426 (2001)
474 M. Ziese

175. Nagaosa, N., Sinova, J., Onoda, S., MacDonald, A.H., Ong, N.P.: Anomalous Hall effect.
Rev. Mod. Phys. 82, 1539 (2010)
176. Fang, Z., Nagaosa, N., Takahashi, K.S., Asamitsu, A., Mathieu, R., Ogasawara, T., Yamada,
H., Kawasaki, M., Tokura, Y., Terakura, K.: The anomalous Hall effect and magnetic
monopoles in momentum space. Science 302, 92 (2003)
177. Chen, Y., Bergman, D.L., Burkov, A.A.: Weyl fermions and the anomalous Hall effect in
metallic ferromagnets. Phys. Rev. B 88, 125110 (2013)
178. Neubauer, A., Pfleiderer, C., Binz, B., Rosch, A., Ritz, R., Niklowitz, P.G., Böni, P.:
Topological Hall effect in the A phase of MnSi. Phys. Rev. Lett. 102, 186602 (2009)
179. Li, Y., Xiong, P., von Molnár, S., Wirth, S., Ohno, Y., Ohno, H.: Hall magnetometry on a
single iron nanoparticle. Appl. Phys. Lett. 80, 4644 (2002)
180. Zeldov, E., Majer, D., Konczykowski, M., Geshkenbein, V.B., Vinokur, V.M., Shtrik-
man, H.: Thermodynamic observation of first-order vortex-lattice melting transition in
Bi2 Sr2 CaCu2 O8 . Nature 375, 373 (1995)
181. Zhang, S.: Spin Hall effect in the presence of spin diffusion. Phys. Rev. Lett. 85, 393 (2000)
182. Dyakonov, M.I., Perel, V.I.: Current-induced spin orientation of electrons in semiconductors.
Phys. Lett. 35A, 459 (1971)
183. Dyakonov, M.I., Khaetskii, A.V.: Spin Hall effect. In: Dyakonov, M.I. (ed.) Spin Physics in
Semiconductors, pp. 211–243. Springer, Berlin/Heidelberg (2008)
184. Hirsch, J.E.: Spin Hall effect. Phys. Rev. Lett. 83, 1834 (1999)
185. Sinova, J., Valenzuela, S.O., Wunderlich, J., Back, C.H., Jungwirth, T.: Spin Hall effects.
Rev. Mod. Phys. 87, 1213 (2015)
186. Kato, Y.K., Myers, R.C., Gossard, A.C., Awschalom, D.D.: Observation of the spin Hall
effect in semiconductors. Science 306, 1910 (2004)
187. Kimura, T., Otani, Y., Sato, T., Takahashi, S., Maekawa, S.: Room-temperature reversible
spin Hall effect. Phys. Rev. Lett. 98, 156601 (2007)
188. Valenzuela, S.O., Tinkham, M.: Direct electronic measurement of the spin Hall effect. Nature
442, 176 (2006)
189. Saitoh, E., Ueda, M., Miyajima, H., Tatara, G.: Conversion of spin current into charge current
at room temperature: inverse spin-Hall effect. Appl. Phys. Lett. 88, 182509 (2006)
190. Miao, B.F., Huang, S.Y., Qu, D., Chien, C.L.: Inverse spin Hall effect in a ferromagnetic
metal. Phys. Rev. Lett. 111, 066602 (2013)
191. Bernevig, B.A., Zhang, S.-C.: Quantum spin Hall effect. Phys. Rev. Lett. 96, 106802 (2006)
192. König, M., Wiedmann, S., Brüne, C., Roth, A., Buhmann, H., Molenkamp, L.W., Qi, X.-L.,
Zhang, S.-C.: Quantum spin Hall insulator state in HgTe quantum wells. Science 318, 766
(2007)
193. Hasan, M.Z., Kane, C.L.: Topological insulators. Rev. Mod. Phys. 82, 3045 (2010)
194. Beach, R.S., Berkowitz, A.E.: Giant magnetic field dependent impedance of amorphous
FeCoSiB wire. Appl. Phys. Lett. 64, 3652 (1994)
195. Panina, L.V., Mohri, K.: Magneto-impedance effect in amorphous wires. Appl. Phys. Lett.
65, 1189 (1994)
196. Phan, M.-H., Peng, H.-X.: Giant magnetoimpedance materials: fundamentals and applica-
tions. Progress Mater. Sci. 53, 323 (2008)
197. Panina, L.V., Mohn, K., Uchyama, T., Noda, M.: Giant magneto-impedance in Co-rich
amorphous wires and films. IEEE Trans. Magn. 31, 1249 (1995)
198. Yoon, S.S., Kim, C.G.: Separation of reversible domain-wall motion and magnetization
rotation components in susceptibility spectra of amorphous magnetic materials. Appl. Phys.
Lett. 78, 3280 (2001)
199. Chen, D.-X., Muñoz, J.L.: Ac impedance and circular permeability of slab and cylinder. IEEE
Trans. Magn. 35, 1906 (1999)
200. Yelon, A., Ménard, D., Britel, M., iureanu, P.: Calculations of giant magnetoimpedance and
of ferromagnetic resonance response are rigorously equivalent. Appl. Phys. Lett. 69, 3084
(1996)
9 Magnetotransport 475

201. Montgomery, H.C.: Method for measuring electrical resistivity of anisotropic materials. J.
Appl. Phys. 42, 2971 (1971)
202. van der Pauw, L.J.: A method of measuring the resistivity and Hall coefficient on lamellae of
arbitrary shape. Philips Tech. Rev. 20, 220 (1959)
203. Büttiker, M.: Symmetry of electrical conduction. IBM J. Res. Develop. 32, 317 (1988)
204. Johnson, M.: Spin accumulation in gold films. Phys. Rev. Lett. 70, 2142 (1993)
Magneto-optics and Laser-Induced
Dynamics of Metallic Thin Films 10
Mark L. M. Lalieu and Bert Koopmans

Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 478
Magneto-Optics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 480
Basics of Magneto-Optics from a Macroscopic Perspective . . . . . . . . . . . . . . . . . . . . . . . . . 480
Micropic Understanding of Magneto-Optics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 485
Measuring Magnetism Using MOKE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 488
Measuring Ultrafast Magnetization Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 497
Ultrafast Laser-Induced Magnetization Dynamics and Opto-Magnetism . . . . . . . . . . . . . . . . 502
Conceptual Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 502
Ultrafast Laser-Induced Loss of Magnetic Order . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 505
All-Optical Switching of Magnetization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 521
Laser-Pulse-Excited Spin Currents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 527
Conclusions and Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 534
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 537

Abstract

Optical tools have played an important role in the field of thin film magnetism,
and the spintronics that has emerged from it. Initially, the role of optics
was limited to providing sensitive and versatile magneto-optical probes of the
magnetic behavior. In the 1990s, it became clear that ultrashort laser pulses can
also be successfully used in a reverse approach to manipulate magnetic order at
down to femtosecond time scales, an approach we refer to as opto-magnetism.
This chapter provides a basic introduction and a review of developments in
both magneto-optics and opto-magnetism, as mostly applied to metallic thin
film ferri- and ferromagnetic systems. In the second part of the chapter, after a

M. L. M. Lalieu · B. Koopmans ()


Department of Applied Physics, Eindhoven University of Technology, Eindhoven,
The Netherlands
e-mail: b.koopmans@tue.nl

© Springer Nature Switzerland AG 2021 477


J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_10
478 M. L. M. Lalieu and B. Koopmans

conceptual introduction of ultrafast laser-induced magnetization dynamics, three


processes are treated in more detail: laser-induced loss of magnetic order, all-
optical switching of magnetization, and pulsed laser-induced spin currents.

Introduction

This chapter discusses how light can be used to unveil magnetic ordering phe-
nomena in thin film systems but also how it has become a unique tool to modify
magnetism at unprecedentedly short time scales. Key to probing magnetism is the
notion that polarized light changes its polarization state when interacting with a
magnetically ordered material. This was discovered by the English scientist Michael
Faraday in 1845. He found that the plane of polarization is rotated when light
propagates through a magnetized material [1]. The effect is nonreciprocal, i.e., it
changes sign upon reversal of the magnetization, it is cumulative in the sense that the
polarization rotation is proportional to the optical path length through the material
under investigation, and it has its origin in the different refractive index experienced
by left-handed and right-handed circularly polarized light. A similar effect occurs
when light is reflected from a magnetized material. However, in reflection it lacks
the cumulative behavior, and therefore the rotation is generally much smaller. For
this reason, it was only discovered 30 years later in 1877 by the Scottish scientist
John Kerr [2]. The effects in transmission and reflection are referred to as the
Faraday effect and the magneto-optical (MO) Kerr effect (MOKE), respectively.
Nowadays, MOKE has become one of the most versatile and widely used
characterization techniques in research on (thin film) ferromagnetic metals. Its
popularity got a strong impulse in the 1980s with the advance of thin film magnetism
that lead to the birth of spintronics.
The complementary approach, using light not to probe magnetism, but to manip-
ulate magnetic order, has a much shorter history. Among the simplest applications
is heating a ferromagnetic thin film by a focused laser spot and thereby reducing the
magnetic moment according to equilibrium thermodynamics. Since the magnetic
anisotropy has an even stronger temperature dependence than the magnetization
itself, this approach can be used for thermomagnetic writing. It was applied e.g.
in the MO recording strategies investigated in the 1990s [3]. Apart from this
technological drive, researchers were challenged by the question as to how rapidly
the magnetization could be quenched upon sudden laser heating.
The pioneering experiments on this matter were conducted by Beaurepaire and
co-workers in 1996 [4]. The authors investigated laser-induced magnetization
dynamics in the strongly nonequilibrium regime using femtosecond laser pulses.
They found that the demagnetization in ferromagnetic nickel thin films was
completed well within a picosecond, much faster than expected; see Fig. 1a. This
observation has become a benchmark in the field. Unfortunately, both the first author
of the seminal paper, Eric Beaurepaire, and his co-author Jean-Yves Bigot passed
away in 2018.
10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films 479

Fig. 1 Pioneering examples of ultrafast opto-magnetism – the reverse of magneto-optics. (a)


Iconic measurement by Beaurepaire et al. showing the measured MO contrast as a function
of pump-probe delay time for a Ni thin film. It provided the first evidence for sub-ps loss of
magnetization after laser heating with a fs laser pulse. Figure taken from Ref. [4]. (b) Artistic
impression of the first ever reported all-optical switching of magnetization, showing alternatively
up (white) and down (gray) magnetic domains written by laser pulses with opposite helicities,
based on a real data set. (Figure taken from Ref. [5] with kind permission from APS)

Fueled by the novel insight, a new field in magnetism emerged. Much effort is
still being devoted to understanding the ultrafast laser-induced demagnetization, but
this phenomenon turned out just to be a tip of an iceberg. Many new mechanisms
to manipulate ferromagnetic matter by short pulses of light have been discovered
since – a field that we will broadly refer to as opto-magnetism.
Among the highlights was the demonstration in 2007 that a single fs laser
pulse can be used to “write” the magnetic moment of ferrimagnetic GdFeCo
deterministically, where the final direction is merely determined by the helicity of
the circularly polarized light [5], Fig. 1b. Soon thereafter, an alternative form of
all-optical switching (AOS) of the magnetization was found, which is independent
of the laser polarization and leads to a toggle-type behavior that reverses the
magnetization at the impact of each laser pulse [6, 7]. Furthermore, while AOS
was originally restricted to ferrimagnetic alloys (materials with spin subsystems
with oppositely oriented magnetizations), very similar results have been found for
synthetic ferrimagnets [8], i.e., multilayered thin films composed of layers with
opposite magnetization and even simple ferromagnetic films with perpendicular
magnetic anisotropy, such as Pt/Co/Pt [9]. In all these cases, however, the AOS
turned out not be the result of a single fs pulse but a cumulative effect needing many
pulses.
In parallel to these discoveries, it was found that spin currents are produced
when ultrashort laser pulses are absorbed in ferromagnetic thin films [10, 11]. The
spin angular momentum carried by these currents can affect the local magnetization
dynamics [10] but can also exert a spin-transfer torque on a nearby, noncollinearly
480 M. L. M. Lalieu and B. Koopmans

aligned magnetic layer [12, 13]. Altogether, the field of opto-magnetism is in a


highly excited phase, where new discoveries are being made, while experiment and
theory are collaborating to unravel the underlying microscopic mechanisms.
This chapter provides an introduction to both magneto-optics and opto-
magnetism, particularly focusing on ferro- and ferrimagnetic metallic thin films.
In section “Magneto-Optics”, we first introduce the basics of magneto-optical
phenomena, both from a phenomenological and microscopic perspective. Then,
we will treat a variety of magneto-optical schemes to study static magnetic order
in (multilayered) thin film systems, followed by time-resolved extensions thereof
employing pulsed pump-probe laser strategies. Specific implementations to obtain
layer- or element-specific information and MO microscopy will be discussed. In
section “Ultrafast Laser-Induced Magnetization Dynamics and Opto-Magnetism”,
we switch to the use of laser pulses to excite magnetization dynamics. After an
introduction of relevant concepts, we focus on three specific contemporary topics:
ultrafast demagnetization, all-optical switching, and laser-induced spin currents. We
conclude with some final remarks and a brief outlook in section “Conclusions and
Outlook”.

Magneto-Optics

In this section, the principles of the magneto-optical Kerr effect are introduced,
both from a phenomenological and a microscopic perspective. Different schemes
for MOKE measurements on metallic thin films are explained, both for static and
ultrafast time-resolved studies. Ways to separate magnetic signal from different
layers or elements, and Kerr microscopy are addressed.

Basics of Magneto-Optics from a Macroscopic Perspective

Magneto-optics describes the interaction of light with magnetized matter, leading


to a change in its intensity or polarization state. In this chapter, we mostly focus
on the application to ferro- and ferrimagnetic metals, which have an optical skin
depth of typically 10–20 nm only. Experiments are thus limited to the reflection
geometry or using thin enough films on top of an optically transparent substrate.
For ultrathin films, the rotation of the polarization is typically of the order of tens
of millidegrees only. Yet, they have been proven to provide a versatile and very
sensitive measure of the magnetization using a variety of experimental schemes. As
an intriguing example, minute laser-induced dynamic effects of less than a percent
of change in the magnetization in sub-nm ferromagnetic films can be routinely
measured with high signal to noise, as will be discussed throughout this chapter. In
such experiments, the sensitivity to the optical polarization is typically 10−7 − 10−8
rad, corresponding to rotating the holder of a 1 cm polarizer by no more than a
nm at its circumference. In this chapter, we will focus on reflection studies, unless
otherwise specified.
10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films 481

Complementary to the refractive effects introduced before, magnetic circular


dichroism (MCD) is an important manifestation of the interaction of light with mag-
netic matter. As an example, performing a Faraday experiment on a semiconductor
magnetized along the propagation direction of the light, and choosing light with a
photon energy above the band gap of the material, left- and right-handed circularly
polarized light will be absorbed in a different way. Similarly, for ferromagnetic
metals such circular dichroism provides a tool for measuring magnetic properties,
provided the optical thickness of the specimen under consideration allows for
transmission experiments.
The fundamental interaction by light and matter is described most generically by
the nonlocal and nonlinear optical susceptibility χ . It relates the induced dielectric
polarization P to the electric field E in a rather informal notation via
 
(r, r  ) · E(r  )dr  + 0 (r, r  , r  ) : E(r  )E(r  )dr  dr  + · · · ,
(1) (2)
P = 0 χ χ
(1)
(n)
where χ is the n-th order optical susceptibility tensor of rank n + 1 and 0 is the
permittivity of free space. Here, P (t) = P e−iωt denotes the induced polarization at
optical frequency ω. In the right-hand side, only (combinations of) electrical fields
E(t) = Ee−iωi t oscillating at (a set of) frequencies (ω1 , ..., ωn ) that contribute
to a polarization at ω = ω1 + ... + ωn are considered. Although in this chapter we
mainly consider the linear response (n = 1), we will briefly address (magnetization-
induced) optical second-harmonic generation (SHG) as a relevant nonlinear optical
technique (n = 2). For completeness, Eq. (1) includes nonlocal optical response
beyond the electric dipole approximation, such as electric quadrupole and magnetic
dipole processes. However, such nonlocal extensions are considered to be beyond
the scope of this chapter.
Only considering the local and linear optical response, the dielectric tensor  of
rank 2 is related to the (linear) optical susceptibility by
 
(1)
ij = 0 δij + χij , (2)

where i and j ∈ {x, y, z} and δij is the Kronecker-delta function. For isotropic
materials, ij = 0 δij . Here,  is the relative permittivity or dielectric constant.
Note that in this convention  is dimensionless, whereas elements of the dielectric
tensor ij have units C2 · N−1 · m−2 . The dielectric constant is related to the complex
refractive index ñ = n + iκ via  = ñ2 , in which n is the refractive index and κ is
the extinction coefficient.
In order to understand the basics of magneto-optics, we consider an optically
isotropic material with uniform magnetization M. Then, the dielectric tensor reads:
⎛ ⎞
xx xy xz
 = ⎝ −xy xx yz ⎠ . (3)
−xz −zy xx
482 M. L. M. Lalieu and B. Koopmans

All these tensor elements are complex parameters, with a real part (representing
an induced dielectric polarization in-phase with the driving field) and an imaginary
part (a polarization 90 degrees out of phase). Furthermore, note that whereas the
diagonal elements ii are even in M, the nonreciprocal, off-diagonal elements
ij transform antisymmetrically under reversal of the magnetization, M ↔ −M.
The latter property provides the “magnetic contrast,” being either a change of the
polarization of the light after reflection or transmission or a change of dichroic
signal, upon reversal of M. The off-diagonal element ij is related to the component
of M parallel to î×jˆ. Thus, for spatially isotropic materials carrying a magnetization
in the z-direction, the dielectric tensor reduces to
⎛ ⎞ ⎛ ⎞
xx xy 0 1 iQ 0
 = ⎝ −xy xx 0 ⎠ = xx ⎝ −iQ 1 0 ⎠ , (4)
0 0 xx 0 0 1

where in the second step, we adopted the convention of introducing a magneto-


optical parameter Q related to the dielectric tensor via xy = ixx Q.
For any (multi)layered material consisting of an arbitrary number of layers, the
MO Kerr and Faraday rotation at a specific angle of incidence are fully specified
by the thickness and dielectric tensor elements of all the layers. The MO response
is most conveniently calculated using a transfer matrix approach [14]. Here, we
restrict ourselves to a few very simple cases. For convenience, we adapt a Jones
matrix formalism [15], in which the light’s electrical polarization is described by a
vector with two components, representing two orthogonal components of E. Unless
otherwise specified, the two components are assumed to represent the s- and p-
polarization, respectively, i.e., field components perpendicular (s, from the German
senkrecht) and parallel (p) to the plan of incidence, as sketched in Fig. 2a. Within the
Jones formalism, each optical operation (reflection, transmission, or propagation)
acting on a plane wave can be represented by a 2 × 2 matrix. The resulting electrical
field after the operation is then just calculated by multiplying the relevant matrix
with the incident field vector.
As a consequence of the complex nature of the dielectric tensor, the
magnetization-induced change of the light’s polarization, referred to as the complex
Kerr rotation θ̃ , is a complex quantity consisting of a real and imaginary part. Let
us consider the simplest case of a polar configuration with light incident at arbitrary
angle to an isotropic thin film sample having an out-of-plane magnetization. The
Jones matrix of this sample can be written as

rss rsp
, (5)
rps rpp

where the diagonal elements represent the ordinary reflection coefficients for s-
and p-polarized light. The off-diagonal elements give rise to the magneto-optical
effects. They are (generally) orders of magnitude smaller than the diagonal ones
10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films 483

polar longitudinal transverse


p
s

(a) (b) (c) (d) (e)


Fig. 2 (a) Definition of s- and p-polarization. (b) Illustration of the MO Kerr effect, in which
linearly polarized light gains a Kerr rotation θ and ellipticity ε upon reflection of a magnetic thin
film, with ε defined as the ratio of the minor to the major axis of the elliptically polarized light.
(c-e) Three configurations of MOKE; (c) polar configuration, (d) longitudinal configuration, and
(e) transversal configuration

and change sign when reversing M. Considering the incident light to be s-polarized
and normalizing all field amplitudes to the incident one, the reflected field can easily
be derived to be


rss rsp 1 1
= rss . (6)
rps rpp 0 rps /rss

Thus, we found that the polarization has rotated by an angle ±θ̃ = ±θ + iε =


rps /rss , where ± refers to a magnetization in the ±ẑ-direction and θ and ε are the
Kerr rotation and ellipticity, respectively; see Fig. 2b. Readers familiar with ellip-
sometry will immediately realize the importance of having two complementary opti-
cal responses for extracting physical information from more complex (multi)layered
systems. In section “Layer-Specific MOKE”, it will become clear how one can profit
from this. It is of relevance to stress that upon p-polarized incidence, the complex
Kerr rotation is given by ±θ̃ = rsp /rpp = rps /rss . This means that, in general, the
MO signal will depend on the chosen polarization of the incident light.
Aiming for a calculation of the (complex) MO Kerr and Faraday rotation, a more
transparent description is obtained in terms of the eigen modes of Eq. (4), which
are circularly polarized electro-magnetic waves. Considering propagation along z,
we can define dielectric elements + and − , experienced by the right- and left
circularly polarized eigen modes, respectively. They are related to xy via

+ − − = ixy . (7)

In this notation, modes with opposite circular polarization experience a different


complex refractive index, ñ± , according to

ixy
Δñ ≡ ñ+ − ñ− = = −ñQ. (8)
0 ñ
484 M. L. M. Lalieu and B. Koopmans

Due to the difference between ñ+ and ñ− and restricting ourselves to real values,
right and left circularly polarized light will experience a relative phase difference
while propagating in a direction collinear with M. Considering linearly polarized
light, which is a linear superposition of circularly polarized light with opposite
handedness,




1 1 1 1 1 1
=√ √ +√ , (9)
0 2 2 i 2 −i

propagating through a lossless medium (ñ = n, κ = 0) over a distance d, one readily


obtains for the transmitted electrical field, neglecting an overall time-dependent part
e−iωt


ein+ ωd/c 1 ein− ωd/c 1 ind cos(nQωd/2c)


+ =e . (10)
2 i 2 −i sin(nQωd/2c)

This result describes a rotation of the initial polarization that is proportional with
d, which is just the Faraday effect. Similarly, it is straightforward to verify that
dissipative media with magnetic circular dichroism, i.e., Δκ = 0, generally show
both a polarization rotation and a change in ellipticity.
Of more relevance in the context of thin film magnetism is the difference in the
reflection coefficient experienced by right and left circularly polarized light, which
for normal incidence on a semi-infinitely thick material magnetized perpendicular
to the surface reads
1 − ñ±
r± = . (11)
1 + ñ±
It can be easily verified that reflected linearly polarized light will experience a
(complex) rotation of its polarization, given by

iñQ
θ̃ = , (12)
−1
which is just the MO Kerr effect at normal incidence. As stated before, at non-
normal incidence, the complex Kerr rotation depends on the polarization of the
incident light. Similar expressions for (multi)layered thin film systems and arbitrary
configurations can readily be derived using the earlier mentioned transfer matrix
method [14].
In the analysis so far, magneto-optics was described in terms of a dielectric
tensor that incorporates the magnetization implicitly. Alternatively, one may lin-
earize the M-dependence and introduce a material-dependent – but magnetization-
independent – tensor α accordingly. This description is particularly illustrative
when, in section “Ultrafast Laser-Induced Magnetization Dynamics and Opto-Mag-
netism”, we will treat the so-called inverse-Faraday effect (IFE). For a spatially
isotropic material with a finite magnetization, the induced dielectric polarization P
can be written as
10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films 485

P = αE × M, (13)

where α is the diagonal element of α, and E is again the amplitude of the oscillating
electric field, and M is the static magnetization vector. Equation (13) provides an
intuitive picture of both the Faraday effect and MOKE. Assuming Mẑ and incident
light with wave vector qM and polarization Ex̂, an induced oscillatory dielectric
polarization P ŷ is generated, emitting electro-magnetic radiation coherent with
the incident light but with a polarization orthogonal to it. Superposition of the
transmitted or reflected incident wave with the orthogonally polarized one is
observed as a rotation of the plane of polarization.
To complete this phenomenological introduction of MO effects, three different
configurations in which MOKE can be applied are introduced. They are distin-
guished by the orientation of the magnetization vector with respect to the sample’s
surface normal and the plane of incidence of the light; see Fig. 2c-e. In the polar
geometry, the magnetization vector is perpendicular to the magnetic thin film (or
sample surface). In the longitudinal configuration, the magnetization is parallel to
the film and lies within the plane of incidence. In this case, a non-normal incidence
of the light needs to be chosen, since a normal incidence would result in qP , for
which the MO effects vanish; see Eq. (13). This is in contrast to the polar Kerr
effect, in which case perpendicular incidence can be considered the most simple
and efficient configuration. Finally, the transverse configuration refers to the case
where the magnetization vector is parallel to the magnetic film and perpendicular
to the plane of incidence. This configuration requires p-polarized light, because the
signal vanishes for pure s-polarized light (Eq. 13). In the transverse configuration,
no change of polarization can be observed, but the Kerr effect manifests itself as a
change in intensity upon reversal of the magnetization. As a rule of thumb, the polar
effect is typically an order of magnitude stronger than the longitudinal effect.

Micropic Understanding of Magneto-Optics

The phenomenological analysis so far was purely based on symmetry. Aiming for a
microscopic understanding, the elements of the dielectric tensor – both the diagonal
and the off-diagonal ones – can be related to the electronic states of the material.
Within the electric dipole approximation, the interaction between light and matter
can be treated equivalently [16] using either an interaction Hamiltonian eE · r
(sometimes referred to as velocity gauge) or −(e/m)A · p (length gauge), where
e is the electron charge, m its mass, A is the vector potential, and r and p = −ih̄∇
are the position and momentum operator, respectively. For isolated atoms this leads
to the well-known dipole selection rule ΔLz = ±1, where Lz is the z-component
of the angular momentum operator, which allows for transitions between s- and
p-states, and between p- and d-states, but not between s- and d-states.
Within a single-particle description, the dielectric tensor elements can be
expanded as a sum over contributions due to “vertical” optical transitions (k n = k n ),
introducing matrix elements
486 M. L. M. Lalieu and B. Koopmans

 = Ψn |pi |Ψn
,
i
Πnn (14)

where |Ψn
are single-particle electronic states with wave vector k n . Following Ref.
[17], converting equations from c.g.s. to S.I. units and expressing elements of the
dielectric tensor ij (ω) = 0 δij + iσij (ω)/ω rather than the optical conductivity
σij (ω), one obtains

e2 f (εn ) − f (εn ) Π i  Π j 
ij (ω) = 0 δij + n n nn
. (15)
h̄m2 V ω  ωnn ω − ωnn + i/τ
n,n

Here, the summation is over all states n (n ) with energy εn (εn ) and h̄ωnn =
εn − εn . Furthermore, f (εn ) refers to the Fermi-Dirac distribution function, and
V is the volume of the crystal. Finite lifetime effects have been taken into account
in a phenomenological way, by introducing a life time τ for all transitions.
In order to capture the magneto-optics in a transparent way, the matrix elements
can be written in an alternative form that reflects transitions by circularly polarized
light, i.e., photons carrying a positive or negative quantum of angular momentum:

±  
Πnn  = Ψn |pi − ipj |Ψn , (16)

for light propagating in the direction î × jˆ. In this notation, the off-diagonal elements
of the dielectric tensor read [17, 18]

 2  j 2
ie2 Πni  n  − Πn n 
occ. unocc.
ij (ω) = , with i = j, (17)
2h̄m2 V ω n n
ωn2 n − ω2

where now the summation over n ranges over all occupied states and the summation
over n over all empty states. The full response including the corresponding real part
can be obtained by Kramers-Kronig transformation [18].
It can be shown that the quantity |Πn+ n |2 − |Πn− n |2 , relevant for the nonreciprocal
off-diagonal elements ij , vanishes once summed over any full set of degenerate
states if spin-orbit coupling is neglected.1 This reflects the fact that magneto-optics
probes (a weighted average of) orbital angular momentum rather than spin ordering.
Thus, although usually it is assumed that magneto-optics is sensitive to the spin
ordering, this is only true in the case of finite spin-orbit coupling and assuming
a strict correlation between spin and orbital momenta. Exact summation rules can
be derived that show that the integrated spectral weight of a certain nonreciprocal
tensor element corresponds to a component of the total orbital momentum vector
[20]. More loosely speaking, one can say that magneto-optics measures the orbital
moment, whereas the spin moment can only be derived by assuming a fixed
relation between the two. This should be kept in mind particularly when performing
10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films 487

experiments on strongly nonequilibrium systems, as will be discussed later in


section “Ultrafast Laser-Induced Loss of Magnetic Order”.
The inability to directly probe the spin moments is reflected in the MO response
of elementary (3d transition-metal) ferromagnets and their alloys. In these materials,
the orbital moments are strongly quenched by crystal-field effects, so their magneti-
zation is strongly dominated by the spin ordering, while orbital moments provide
only a small contribution. As a consequence, MO effects are relatively small.
Examples of experimental MO spectra for the elementary ferromagnets nickel,
cobalt, and iron, displaying xy as a function of photon energy, are depicted in Fig. 3.
Features such as peaks in the spectrum can be assigned to interband transitions.
Together with the zero crossings, they lead to the notion that the magnitude of the
Kerr effect for a specific material and even its sign depend strongly on the photon
energy. A simple intuitive picture explaining peaks in xy for a transition-metal
ferromagnet is sketched in Fig. 4. Panel (a) shows a simplified density of states
for the d-bands of spin-up and spin-down electrons, split by an exchange splitting
of several tenths of eV. The bands further split up by a weaker spin-orbit coupling,
described by

a 0.8 b 4

0.6 Im 3
Im
0.4
(eV 2)

(eV 2)

2
0.2
0

0.0 1
/

/
xy

xy

Re
(h )2

(h )2

-0.2
0
-0.4
-1 Re
-0.6

-0.8 -2
0 1 2 3 4 5 6 0 1 2 3 4 5 6
h (eV) h (eV)

c 4

3 Im
(eV 2)

2
0

1
/ xy
(h )2

-1
Re
-2
0 1 2 3 4 5 6
h (eV)

Fig. 3 Experimental MO spectra showing the square of the real (black) and imaginary (red) value
of (h̄ω)2 xy /0 for (a) nickel, (b) cobalt, and (c) iron, as a function of the photon energy h̄ω. The
data points are adopted from Ref. [21], and the solid lines are guides for the eye
488 M. L. M. Lalieu and B. Koopmans

a b

Fig. 4 (a) Simplified spin-resolved density of states (up and down electrons), showing exchange
and spin-orbit split d-bands and one of the p-states out of the broad sp-band. (b) Peaks in the
imaginary part of the diagonal dielectric tensor element xx for spin-conserving transitions between
p and d states for up and down electrons, as derived from (a). (c) Similar construction of the
imaginary part of xy . The real parts, both in (b) and (c), can be obtained by Kramers-Kronig
transformations. (Figure adapted from Refs. [22, 23])

Hso = ξ L · S, (18)

where ξ is the material-dependent spin-orbit coupling parameter. The splitting into


Lz = −2 · · · + 2 sublevels is opposite for the up and down bands, as apparent from
Eq. (18). Considering optical transitions between a single (degenerate) p-state out of
the broad sp-band and the band of nondegenerate d-states and taking into account
the dipole selection rule ΔLz = ±1 for right and left circularly polarized light,
respectively, one readily derives the schematic spectra for the absorptive (imaginary)
parts of the diagonal (xx ) and non-diagonal (+ − − = ixy ) dielectric tensor
elements; see Fig. 4b, c, respectively.
When interfacing the transition metals with heavy elements like Pt or Pd, spin-
orbit effects can be enhanced. This may lead to larger orbital moments with
enhancement of the MO Kerr effect and other consequences such as perpendicular
magnetic anisotropy. As an example, experimental values for the Kerr rotation of the
polarization axis as observed in Co/Pt multilayers range typically up to 0.5◦ [24].

Measuring Magnetism Using MOKE

Different configurations for performing MOKE experiments will be briefly intro-


duced, starting with a simple crossed-polarizer approach and followed by more
sensitive schemes using a balanced photodiode and polarization modulation. Sub-
sequent sections capture modes of operation that further enhance the applicability,
addressing layer-specific studies, spectroscopic MOKE, Kerr microscopy, and other
advanced approaches.
10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films 489

Different Configurations
We consider a reflection experiment with linearly polarized light, incident on a flat
surface at a certain angle of incidence. The polarization state of the light will be
expressed as an angle θin , according to tan(θin ) = Es /Ep , where s and p refer
to the s- and p-polarized components (Fig. 2a). Upon reflection, the polarization
state may be changed, and we denote the change of rotation by θ ≡ θout − θin ,
where “in” and “out” refer to the incoming and reflected light. It should be
realized that a change of polarization state can already be induced without any
MO interaction and even from an isotropic material. This generally occurs when
the incident light is neither s- nor p-polarized. The resulting effects thereof are
widely used in ellipsometry. For the analysis of MO measurements, it is convenient
to distinguish between contributions to the total (complex) polarization rotation
that transfer symmetrically (“S”) and antisymmetrically (“A”) under reversal of the
magnetization M, θ̃T = θ̃S + θ̃A . Then, the complex Kerr rotation θ̃ = θ + i as
introduced in section “Basics of Magneto-Optics from a Macroscopic Perspective”
is obtained from

θ̃T (M) − θ̃T (−M)


θ̃ = (19)
2

– a common experimental procedure to extract magnetic information in


polarization-sensitive studies. In this manner, as will be demonstrated later
(section “Measuring Ultrafast Magnetization Dynamics”), any nonmagnetic
contribution to the optical signal in a MOKE measurement can be elimi-
nated.
In the generic case of a sample with a depth profile of the magnetization, and
possibly having different magnetic sublattices that are labeled with an index i, the
magnetization is described by M i (z) (with z ≥ 0 and as measured with respect to
its surface at z = 0). For such a system, we can relate the measured (complex) Kerr
rotation to the magnetization profile via

 ∞
θ̃ = F̃ i (z) · M i (z)dz, (20)
i 0

where F̃ i (z) is a vector of generalized Fresnel coefficients, depending on all details


of the experimental configuration, the wavelength of light being used, and the
sample layout. F̃ i (z) determines the sensitivity to the magnetization of sublattice
i at depth z and will generally decay at a length scale corresponding to the optical
skin depth in the material. The simplest case occurs when (1) one measures a single
magnetic film with uniform magnetization M along a known direction or (2) having
chosen the configuration or sample layout such that all components of F̃ (z) are
approximately zero, except for those of a single layer (and single sublattice) with
uniform M i along a known direction. In that case, M i can be taken out of the
integral, and the generic relation reduces to
490 M. L. M. Lalieu and B. Koopmans

θ̃ = F̃ M. (21)

This provides two complementary channels to measure the same magnetic parame-
ter M, i.e., the real and imaginary part of θ̃ = θ +iε. Clearly, in order to optimize the
signal-to-noise ratio, the channel with the largest component of F should be chosen.
Moreover, optical components, such as wave plates, can be used to effectively
measure an optimal linear combination of θ and ε.
Having two complementary channels can be exploited in many ways. First of
all, in the simple case described by Eq. (21), the complementary channels can be
used to verify the assumptions that are made. For example, if reversing the in-
plane magnetization shows up differently in rotation and ellipticity in a longitudinal
MOKE experiment, it could be a signature of a perpendicular component being
mixed in due to a small misalignment of the applied magnetic field. More impor-
tantly, however, the complementary channels can be used to separately measure
different magnetic components, such as the switching of two different magnetic
layers (see section “Layer-Specific MOKE”), to differentiate between an out-of-
plane and in-plane component of the magnetization or to distinguish two different
magnetic sublattices in an alloy.
The simplest realization of a MOKE experiment exploits a crossed polarizer
configuration, as sketched in Fig. 5a. Usually, a collimated light source (often a
CW or pulsed laser) is polarized (P1) before being focused on the sample, the latter
to establish spatial resolution. The reflected light is collimated and sent through a
polarizer (P2), dubbed the analyzer, set at an angle α nearly – but not exactly –
crossed to the polarization axis of P1, i.e., α = |αP1 − αP2 | = π/2 + γ , for small
γ . For simplicity, we assume P1 to be aligned along the s- or p-polarization axis. In
that case, the intensity after the analyzer is given by

I = I0 R(γ 2 + 2γ θ + θ 2 + ε2 ), (22)

where I0 is the intensity of the incident light, R = rr ∗ is the reflection coefficient


for the intensity, and further reflection and absorption losses of optical components

a b c

Fig. 5 Illustration of three different schemes for performing MOKE. (a) A crossed-polarizer
configuration, in which an analyzer (P2) is almost crossed with a polarizer (P1). (b) A balanced
photodiode configuration, in which a polarizing beam splitter (PBS) is adjusted so as to balance
signals of two orthogonal polarization components. (c) A polarization modulation configuration, in
which a PEM is used to establish high-frequency polarization modulation, and a lock-in amplifier
detects signals at the fundamental and second-harmonic frequency of the modulation. A quarter-
wave plate (QWP) can be inserted to select a linear combination of rotation and ellipticity
10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films 491

are neglected. The terms quadratic in θ and ε can usually be neglected because of
the smallness of the Kerr effect and, moreover, are symmetric under reversal of M
and thereby are difficult to distinguish from the first term. The 2γ θ term is the Kerr
rotation signal of interest. While it looks as if the sensitivity increases upon opening
the analyzer (i.e., increasing γ ), this comes at the cost of an increased background
signal, described by the first, γ 2 , term. Thus, a compromise should generally be
sought to optimize signal to noise. Finally, note that although according to Eq. (22)
the crossed-polarizer configuration seems to be insensitive to the Kerr ellipticity,
adding a quarter-wave plate (at an angle of 45◦ ) between the sample and the analyzer
reverses the role of rotation and ellipticity.
In an alternative approach, a polarizing beam splitter (PBS) is adjusted to
balance the signal from two photodiodes, measuring orthogonal polarization com-
ponents, as sketched in Fig. 5b. In this balanced photodiode scheme, with the
beam splitter set such that it splits components with polarization angles of +45◦
and −45◦ , the measured intensity simply reads (neglecting terms quadratic in
θ and ε)

I = I0 Rθ. (23)

In comparison with the crossed-polarizer configuration, where noise in the intensity


of the light source enters via the γ 2 term, the balanced photodetector approach is – in
lowest order – just limited by electronic noise and noise related to the light detection.
Another way of enhancing signal-to-noise is by using polarization modulation, as
depicted in Fig. 5c. A popular version uses a high-frequency photo-elastic modulator
(PEM), typically operating at f = 50 kHz. The PEM is installed before the sample,
and the incident light has its linear polarization axis making an angle of 45◦ with
the main axis of the modulator. The Jones matrix of the PEM is given by

1 0
, (24)
0 eiA0 cos(2πf t)

where A0 determines the amplitude of the oscillatory retardation between the two
orthogonal polarization components of the light. Setting A0 = π/2, the light after
the PEM oscillates between left- and right-handed circular polarization. Two simple
configurations can be distinguished. For the configuration where the main axis of
the modulator is set to 45◦ with respect to the plane of incidence of the light on
the sample, we refer to Refs. [25, 26]. Here, we will briefly discuss the simpler case
where the main axis of the modulator is set to 0◦ . The signal from the photodetector,
positioned after the analyzer that is aligned with the PEM at the p-polarization axis,
is sent to a lock-in amplifier. It can be shown [26] that dc and harmonic signals are
generated. Up to second order in the polarization modulation frequency f , they read

Idc = [1/2 + J0 (A0 )θ ] RI0 ≈ RI0 /2, (25)


I1f = J1 (A0 )εRI0 , (26)
I2f = J2 (A0 )θ RI0 , (27)
492 M. L. M. Lalieu and B. Koopmans

where Jn (A0 ) is the n-th order Bessel function. Thus, it is seen that ellipticity ε (via
the 1f -signal) and rotation θ (via the 2f -signal) – or, alternatively, complementary
linear combinations thereof when using an additional wave plate – can be measured
simultaneously.

Layer-Specific MOKE
In case of a multilayer structure including multiple magnetic layers, the measured
value of ε or θ is proportional to a weighted average of the magnetization in
the different layers, where the weighting factors depend on the details of the
structure (materials, layer thicknesses) and measurement configuration (wavelength,
polarization, and angle of incidence), c.f., Eq. (20). In this section, an extension to
the previously discussed MOKE setup will be introduced, which makes it possible to
perform layer-specific measurements in a magnetic bilayer system, i.e., individually
measuring the magnetization of one of the two magnetic layers [27, 28]. The layer-
specific technique discussed here is based on the PEM configuration of the MOKE
setup, but it is noted that it works for other MOKE configurations as well.
A phenomenological explanation of the layer-specific MOKE measurement
technique starts by representing the MO response of a single magnetic layer by a
Kerr vector θ̃ = θ + iε = Ωeiξ in the complex Kerr plane spanned by ε and θ .
The Kerr vector is described by a Kerr amplitude Ω and Kerr angle ξ , as illustrated
in Fig. 6a. Within this representation, the I1f (I2f ) signal in the ordinary MOKE
setup measures the projection of the Kerr vector on the imaginary (real) axis. In
the case of a magnetic bilayer, the Kerr vector describes the MO signal of the
complete structure and is equal to the sum of the Kerr vectors of the individual
layers, θ̃tot = θ̃1 + θ̃2 = F˜1 M1 + F˜2 M2 , as illustrated in Fig. 6b, and where we
again simplified Eq. (20). Measuring, for instance, I2f (i.e., θ ), yields the projection

a b c

Fig. 6 (a) The Kerr vector in the complex plane spanned by the Kerr ellipticity ε and the Kerr
rotation θ. It is described by a Kerr amplitude Ω and Kerr angle ξ , and represents the MO response
of a single magnetic layer. The measured values of ε or θ in the MOKE setup are the projections of
the Kerr vector on the imaginary or real axes. (b) Total Kerr vector representing the MO response
of a magnetic bilayer. The total Kerr vector θ̃tot is equal to the sum of the Kerr vectors of the
individual layers, θ̃1 and θ̃2 . Rotating the projection axis P to be perpendicular to θ̃1 causes the
MO signal of this layer to be excluded in the projection. (c) Kerr vector for different depths di of
a magnetic layer within a structure, with d1 < d2 < d3 . Burying the magnetic layer deeper into
the structure causes a rotation of the Kerr vector, as well as a decrease in its amplitude. (Figures
adapted from Ref. [27])
10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films 493

of θ̃tot on the real axis, also called the projection axis P , and thus includes the MO
signal of both layers. Measurement of an individual layer would be possible when
the projection axis P is rotated to be perpendicular to one of the two Kerr vectors,
e.g., θ̃1 , as illustrated in Fig. 6b. In that case, the projection of θ̃1 is zero, and the
measured signal only includes the projection of θ̃2 . This shows that by rotating the
projection axis, it is possible to individually measure the magnetization in one of
the two layers in a magnetic bilayer by removing the contribution of the other layer
to the measured MO signal.
One possible method to rotate the projection axis is by adding a quarter-wave
plate (QWP) to the MOKE setup [28]. As indicated in Fig. 5c, the QWP is placed
in between the first polarizer and the PEM. Depending on the angle of the QWP,
I1f and I2f measure linear combinations of ε and θ , corresponding to a rotation
of the projection axis. Using Jones matrix calculations [15], the first- and second-
harmonic contributions to the measured intensity of the light can be calculated to be
 
I1f = 2I0 R θ sin(2α) + ε cos2 (2α) J1 (A0 )
  (28)
I2f = 2I0 R θ cos2 (2α) − ε sin(2α) J2 (A0 ),

in which α is the angle of the QWP. It can be seen that for both I1f and I2f , there
are certain angles of the QWP α = α0 at which the MO signal disappears (i.e.,
I1f/2f (α0 ) = 0).
When the Kerr vectors of the two magnetic layers (i = 1, 2) in a magnetic bilayer
have different Kerr angles ξi , as is the case in Fig. 6b, the MO signal of the individual
layers will disappear at different angles α0,i . When the QWP angle is set to α0,i
of layer i, only the magnetization of the other layer is measured, resulting in a
layer-specific measurement. One way to assure different Kerr angles is by using
two magnetic layers made of different materials. This was used in Ref. [28], in
which the magnetization dynamics of the individual Ni and Fe layers in a Fe/Ru/Ni
multilayer were measured separately at different angles of the QWP (α0,Ni = α0,Fe ).
An additional advantage of this technique over other element-specific measure-
ment techniques is the ability to perform layer-specific measurements on magnetic
bilayers containing two identical magnetic layers. This is made possible by a depth
dependence of the MO signal [27, 29], as illustrated in Fig. 6c. The figure illustrates
the Kerr vector of a (single) magnetic layer for different depths di within a structure,
with d1 < d2 < d3 . Burying the magnetic layer deeper into the structure causes a
rotation of the Kerr vector, as well as a decrease in its amplitude. The rotation of
the Kerr vector with depth causes the Kerr angle of two identical magnetic layers in
a magnetic bilayer to be (slightly) different, resulting in a different α0 for the two
layers and making it possible to perform a layer-specific measurement.
An experimental demonstration of the depth dependence of α0 for a single out-
of-plane magnetized Co layer buried in Pt is presented in Fig. 7a. In this figure,
the I2f signal coming from the magnetization in the Co layer is measured as a
function of the QWP angle, for different thicknesses of the Pt capping layer. The I2f
494 M. L. M. Lalieu and B. Koopmans

b 1.0
0.5
0.0
MOKE Signal (norm.)

-0.5
-1.0
1.0
0.5
0.0
-0.5
-1.0
1.0
0.5
0.0
-0.5
-1.0
-250 -200 -150 -100 -50 0 50 100 150 200 250
Field (mT)

Fig. 7 (a) Measurement of the I2f signal coming from the magnetization in the Co layer
as a function of the QWP angle, for different thicknesses of the Pt capping layer. The inset
highlights the shift in α0 when the Co layer is buried under an increasing amount of Pt. The
solid lines are fits to the data using Eq. (28). (b) Layer-specific measurement performed on a
Ta/Pt/Co/Pt/Ru/Pt/Co/AlOx (bottom to top) multilayer, showing three hysteresis loops measured
at three different angles α of the QWP. The top panel shows the hysteresis curve in which both
Co layers are visible, in which the outer (inner) switches correspond to the magnetization in the
bottom (top) Co layer. The middle and bottom curves show layer-specific measurements, in which
the signal of one of the two layers is excluded from the measurement
10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films 495

signal is determined from the height of the hysteresis curve measured using polar
MOKE. The solid lines are fits using Eq. (28), showing a very good agreement.
It can be seen that for each curve there are two angles of the QWP at which the
I2f signal is zero. These angles correspond to the α0 angles of the Co layer. More
importantly, highlighted in the inset, a clear shift in α0 of about 0.4 ◦ nm−1 is seen
when the Co layer is buried under an increasing amount of Pt. This shift correlates
to the rotation of the Kerr vector, as illustrated in Fig. 6c, whereas the decrease in
the signal amplitude with increasing Pt thickness corresponds to the decrease of the
Kerr amplitude.
The prediction that the small shift in α0 allows for a layer-specific measure-
ment in a magnetic bilayer containing two magnetic layers made from the same
material is demonstrated in Fig. 7b. The presented measurement is performed on
a Ta(2)/Pt(10)/Co(1)/Pt(0.4)/Ru(1)/Pt(0.4)/Co(1.3)/AlOx (2.0) (bottom to top and
thickness in nanometer) multilayer, containing two out-of-plane magnetized Co
layers that are antiferromagnetically coupled by a thin Pt/Ru/Pt spacer. At α = 3.57◦
(middle panel), a sensitivity to only the top layer is achieved, resulting in a single-
component hysteresis curve. Similarly, at α = 5.01◦ (bottom panel), only the
bottom layer is seen, while at arbitrary angle (top panel), both layers show up.
This measurement clearly demonstrates that although the difference in α0 for the
two Co layers is quite small (1.44◦ ), a layer-specific measurement is possible,
even when the two magnetic layers are made of the same material, which is not
possible with element-specific measurement techniques such as X-ray magnetic
circular dichroism, discussed in the next section. For completeness, it is noted that
the difference in α0 for the two layers may not only be due to their different depth
but may also be partially ascribed to subtle differences in their electronic and optical
properties due to different neighboring layers.

MO Spectroscopy
Another approach to disentangle magnetic information in more complex systems is
provided by selecting appropriate wavelengths. While most MOKE setups operate
with a single laser or LED operating in the visible or near-infrared, a complete MO
spectrum can be measured using either a broadband light source in combination
with a spectrometer or a wavelength-tunable laser system. Clearly, as with ordinary
ellipsometry, a more complete deconvolution of the MO response in multilayer
systems is possible from such a spectrum, provided enough information about the
wavelength-dependent diagonal and off-diagonal elements of the dielectric tensor
of the different layers is known. More specifically, in some alloys it is known
that different elements contribute to different parts of the MO spectrum. As an
example, in rare-earth transition-metal ferrimagnetic alloys, it is known that the
high-wavelength part (600 nm) is most sensitive to the transition-metal atoms
(and their magnetic moments) and the (ultra)violet part to the rare-earth atoms.
This has been used e.g. to disentangle the different response of the sublattices after
femtosecond laser heating [30].
Extending our scope to sources well outside the visible range, X-ray techniques
are known to be extremely powerful. In fact, X-ray magnetic circular dichroism
496 M. L. M. Lalieu and B. Koopmans

(XMCD) is basically just a special type of MO measurement, though with the


added value of being element-specific by probing the core levels. Moreover, in
special cases, XMCD provides a quantitative separation between spin and orbital
contributions to the magnetic moments. This separation is based on specific
sum rules, as derived, for instance, for the transition-metal L-edges [31, 32].
Interestingly, table-top versions of core level spectroscopy based on high-harmonic
generation (HHG) of fs laser pulses to produce XUV pulses with photon energies
up to ∼100 eV have become available [33]. They can be thought to bridge the
gap between simple MOKE setups and the much more demanding user-facility
synchrotron-based techniques. Nowadays, several groups are performing pump-
probe MO measurements using the HHG pulses as a probe. In contrast to the
fs-XMCD measurements, the approach has been limited to the M-edge of 3d
elements rather than the L-edge and the use of magnetic linear dichroism (MLD).
This development is of particular interest for studies of ultrafast magnetic processes,
e.g., see Ref. [34].
Some examples of full element-specific and spin-orbit-resolved XMCD as well
as table-top core level spectroscopy applied to the exploration of fs magnetization
dynamics will be discussed in section “Ultrafast Laser-Induced Magnetization
Dynamics and Opto-Magnetism”.
Finally, in the context of MO spectroscopy, we briefly address options pro-
vided by higher-order nonlinear terms in Eq. (1), which offer a complementary
optical view of magnetism. Most relevant in the present context is the process
of magnetization-induced second-harmonic generation (MSHG) [35], described
(2)
by the second-order optical susceptibility tensor χ . In this process, occurring
at high laser fluences, light is generated at twice the frequency of the incident
light. Based on symmetry arguments, it can be shown that all tensor elements
vanish in the bulk of centrosymmetric media. It is from this that SHG derives
its intrinsic surface/interface sensitivity. In the late 1980s, it was realized that
the magnetization in a centrosymmetric ferromagnetic metal does break time but
not spatial inversion symmetry. As a consequence, MSHG provides an optical
measurement scheme that is – within the electric dipole approximation – inherently
sensitive to the magnetization at the interface. Absolute MSHG signals are many
orders of magnitude smaller than their linear counterparts. However, this can be
compensated for by the fact that the relative MO contrast can be huge, and in specific
configurations, it can be tuned at will up to 100 % [36]. Some examples of MSHG
as applied to investigations of ultrafast magnetization dynamics will be discussed in
section “Ultrafast Laser-Induced Magnetization Dynamics and Opto-Magnetism”.

MOKE Microscopy
One of the advantages of using MOKE as a magnetometry tool is its spatial
resolution. Even with long-focal-length lenses, a resolution of a couple of tens of
micrometer resolution is easily achieved in a very simple setup. Such resolution
allows local magnetization measurements on thin film samples where one (or more)
of the layers has been given a position-dependent thickness via a “wedge growth”
10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films 497

technique, in which a mechanical knife-edge shutter is withdrawn during deposition


of the respective layer. Additionally, local measurements of MOKE can be easily
implemented in a UHV environment. In the context of ultrathin films and magnetic
surfaces, it has been dubbed surface magneto-optic Kerr effect (SMOKE) [37].
Aiming for a higher resolution, toward the diffraction limit and beyond, a number
of approaches have been developed. In scanning Kerr microscopy a tightly focused
laser spot is used for imaging. The spatial resolution is limited by the Rayleigh
criterion:
λ
s= , (29)
2NA

where s is the smallest distance between features that can be distinguished, λ is


the used wavelength, and NA is the numerical aperture (defined as n sin(θ ) with
θ the collection angle of the microscope objective and n the refractive index of
the medium). However, this approach is not used in practice for separating nearby
magnetic features but for measuring the magnetization of nano-objects significantly
smaller than the diffraction limit. This approach is made possible by the superior
sensitivity of MOKE [38] and was later dubbed “nano-MOKE.” In principle, it can
be implemented using any of the MOKE detection schemes discussed previously.
A complementary, and widely used, approach is wide-field Kerr microscopy, see
Fig. 8a. In fact, this approach is similar to polarization microscopy. It generally
makes use of an intense white light source for illumination, a high-NA microscope
objective, and a crossed-polarizer detection scheme (Fig. 5a). The resolution can
be further enhanced by using an immersion lens (n > 1). In the simplest,
polar, geometry, the full entrance aperture is used. In that case, the incident light
propagates on average perpendicular to the sample’s surface, which leads to a
sensitivity to the perpendicular component of M. In order to gain sensitivity to
in-plane components, light from an asymmetrical part of the entrance aperture
should be selected. When doing this with a four-quadrant detector and measuring
different combinations of signals from the four quadrants, full vectorial resolution
can be achieved; see Fig. 8b. Wide-field Kerr microscopy has been used for decades
to measure micromagnetic domains and their dynamics. Moreover, it has been
extended to the regime of GHz dynamics, often based on stroboscopic approaches
with a variable delay between trains of electronically generated magnetic field
pulses to excite the dynamics and laser pulses for MO probing [41, 42].
Extending the domain of optics further, toward the use of X-rays or photo-
electron spectroscopies, fascinating developments have been reported in the past two
decades. For more details on those novel microscopy techniques, see  Chap. 24,
“Magnetic Imaging and Microscopy” of this handbook.

Measuring Ultrafast Magnetization Dynamics

MOKE provides a very powerful way to measure magnetization dynamics at


ultrafast time scales using fs laser pulses. This is most easily done in a pump-
498 M. L. M. Lalieu and B. Koopmans

a
~16 fps
CCD camera

Longitudinal
Polar Longitudinal with transverse Transverse Transverse
sensitivity
Tube lens

Analyzer
Collector Field
Aperture Compensator
Diaphragm
diaphragm
Xe arc (centered) Polarizer
lamp

Back
focal plane

Objective lens

sample sample

50 mm

Fig. 8 (a) Schematic drawing of a wide-field Kerr microscope. Figure adapted from Ref. [39]. (b)
Example of Kerr microscopy, showing a domain pattern in a permalloy patterned film element
of 240 nm thickness, imaged in longitudinal sensitivities with light incidence from different
directions. (Figure adapted from Ref. [40] Reproduced from [40], with the permission of AIP
Publishing)
10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films 499

probe scheme using two laser pulses, i.e., the first pulse is used to excite (pump) the
magnetization dynamics and the second (mechanically time-delayed) pulse is used
to probe it. The ultimate time resolution is achieved by using two pulses derived
from the same laser source to avoid electronic jitter. To obtain a high sensitivity, a
stroboscopic scheme can be employed, in which a train of high-repetition-rate laser
pulses is focused on the sample. In such cases, the approach is commonly referred
to as time-resolved MOKE (TR-MOKE).
Linearizing Eq. (21) for a simple system in which the MO signal depends solely
on a single magnetization component, it is simple to see that the pump-pulse-
induced complex Kerr rotation can be written as

Δθ̃(t) = F0 ΔM(t) + M0 ΔF (t), (30)

where F0 and M0 are the generalized Fresnel coefficient and the magnetization in the
absence of a pump pulse. This equation carries an important warning. The first term
on the right-hand side is proportional to the laser-induced change in magnetization,
ΔM(t), and is the targeted signal reflecting the genuine magnetization dynamics.
The second term, however, does not. It is due to a pump-induced modification of
the proportionality factor, ΔF (t). We will return to this issue in section “Ultrafast
Laser-Induced Loss of Magnetic Order”, but the message here is that one should
always be aware of the occurrence of possible “artifacts.”
An implementation of TR-MOKE based on the polarization modulation scheme
is sketched in Fig. 9 and will be explained later. A simpler realization exploits the
crossed-polarizer configuration, which was introduced in Fig. 5a. Both the pump
and probe pulses can be derived from the same laser source using a beam splitter
(BS) or frequency converter by a nonlinear crystal. Collimated probe pulses are
polarized before being passed along a mechanical delay line to adjust the time delay
between the two pulses, whereafter they are directed to the sample. Pump and probe
pulses are focused to (at least partially) overlapping spots on the sample. It can be
of advantage to use a probe spot that is smaller than the pump spot, in order to
probe a region where the change in magnetization is more or less homogeneous.
The influence of the pump beam on the polarization state of the reflected probe
pulse is measured using an analyzer and any type of photodetector. To measure
the pump-induced changes to the magnetization, measurements of θ̃(t) with and
without pump-pulse excitation are compared. Often, to enhance the sensitivity, a
mechanical chopper is placed in the pump beam that periodically blocks the pump
pulses (see Fig. 9), and a lock-in amplifier is used to directly measure Δθ̃ (t).
From Eq. (22), it is easy to derive that the pump-induced change in output signal
is described in lowest order of θ̃ by [26]:

ΔI (t)/I0 = 2γ Δθ (t)R0 + 2γ θ0 ΔR(t) + γ 2 ΔR(t), (31)

where R0 and ΔR(t) are the reflectivity and pump-induced transient thereof. Care
has to be taken to rule out artificial signals due to a ΔR(t) of nonmagnetic origin.
The term γ 2 ΔR(t) transforms symmetrically under reversal of M, and can as such
500 M. L. M. Lalieu and B. Koopmans

Fig. 9 Schematic overview of the TR-MOKE setup in a polarization modulation scheme. Using
a delay line to alter the path length of the probe pulse, the magnetization can be measured as a
function of time after the pump-pulse excitation by adjusting the pump-probe time delay (Δt)

be easily removed using a procedure corresponding to Eq. (19). In contrast, the


term 2γ θ0 ΔR(t) transforms antisymmetrically and cannot easily be distinguished
from magnetization dynamics. Bigot et al. argued that part of this drawback of the
crossed-polarizer approach is avoided by performing measurements at a multitude
of analyzer angles [43].
Also the second configuration to measure MOKE, as introduced in Fig. 5b, can be
easily adapted for time-resolved studies [44]. When working exactly at the balanced
configuration, a dependency on ΔR(t) can be avoided [26], and Eq. (23) yields

ΔI (t)/I0 = 2Δθ (t)R0 . (32)

Note that when required, a sensitivity to the complementary ellipticity channel is


obtained by using a quarter-wave plate, an option also available for the crossed-
polarizer configuration.
Finally, the polarization modulation scheme can be applied, as is illustrated in
Fig. 9 (see also Fig. 5c). Using the configuration that was used to derive Eqs. (25)–
(27) and choosing A0 = 2.405, for which J0 (A0 ) = 0, one finds [26]

ΔI1f (t)
= 2J1 (A0 )Δε(t), (33)
Idc
ΔI2f (t)
= 2J2 (A0 )Δθ (t). (34)
Idc

Similar to the balanced photodiode approach, no artifacts proportional to ΔR(t)


appear in lowest order. However, small misalignments can easily give rise to
nonmagnetic contributions. For that reason, it is always recommended to measure
the dynamic response at two opposite orientations of the magnetization whenever
possible. Then, Eq. (19) can be used to deduce the magnetization dynamics.
An example of such a measurement is shown in Fig. 10. In this experiment, the
magnetization in a 10 nm thick Ni film was measured as a function of the pump-
10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films 501

Fig. 10 Typical TR-MOKE measurement performed on an in-plane magnetized 20 nm thick Ni


film, in the presence of an external field applied almost perpendicular to the film surface. Top panel:
time-resolved Kerr rotation measured with the two opposite magnetization directions. Bottom
panel: magnetic (red) and nonmagnetic (blue) contributions to the measured signal, calculated
using the difference and sum of the data measured for the opposite magnetization directions,
respectively. The cartoons illustrate four stages during the observed dynamics; (I) equilibrium state
before laser-pulse excitation with the magnetization aligned to the effective field (red dotted line);
(II) laser-pulse excitation, causing demagnetization and a change in the anisotropy and thereby the
effective field; (III) recovery of the magnetization; (IV) magnetization precession triggered by the
sudden change in effective field direction.( Figure taken from Ref. [45])

probe time delay. The figure displays both the fast demagnetization dynamics in the
first few picoseconds and a laser-induced magnetization precession at a longer time
scale, which is discussed in the following section. The top panel shows the dynamics
measured for the two opposite magnetization directions (black), set by the external
field applied almost perpendicular to the sample surface (H = ±160 kA/m).
The pure magnetic response to the laser-pulse excitation is calculated using the
difference of the two curves (Eq. 19) and is shown by the open dots in the bottom
panel (red). The solid symbols (blue) represent the laser-pulse-induced changes in
the optical signal due to “nonmagnetic” contributions, which is calculated using
the sum of the measured data for the two opposite magnetization directions and
arises e.g. due to state-filling that can occur during pump-probe overlap (see
section “Experimental Demonstration of Laser-Induced Demagnetization”). In this
specific measurement, it only shows some small features around zero pump-probe
delay, but in other cases these nonmagnetic contributions to the signal can become
more prominent, or even dominant.
The setup as displayed in Fig. 9 can be considered as a basic or standard TR-
MOKE arrangement. A simple extension can be made in order to reduce the
nonmagnetic contributions by using a two-color scheme, in which the pump and
502 M. L. M. Lalieu and B. Koopmans

probe pulses have different wavelengths. This can be accomplished, for instance, by
sending the probe beam through a frequency doubler. Also, the previously discussed
layer-specific MOKE technique can be used by adding a QWP to the setup, as was
shown for the static MOKE in Fig. 5c and which allows to individually measure
the magnetization dynamics in one of the two layers in a magnetic bilayer, e.g., see
Refs. [28, 46].
Analogous to the TR-MOKE setup, time-resolved MCD measurements
can be performed by using fs X-ray or XUV pulses for the probe beam,
allowing element-specific and spin-orbit-resolved MCD measurements of the
laser-induced magnetization dynamics. An example of such time-resolved element-
specific XMCD measurement, performed to investigate the laser-pulse-induced
magnetization reversal in a GdFeCo alloy, is discussed in section “All-Optical
Switching of Magnetization”. Lastly, time-resolved MSHG measurements can be
achieved by recording the probe-pulse-induced second-harmonic generation as a
function of the pump-probe delay time [47].

Ultrafast Laser-Induced Magnetization Dynamics and


Opto-Magnetism

Having discussed the use of magneto-optics to probe magnetic phenomena, we now


address the possibility of manipulating the magnetic state by optics – referred to
as opto-magnetism. After a conceptual introduction to the field, three contemporary
topics will be addressed in more detail: (i) the ultrafast loss of magnetic order, so-
called femtosecond (fs) laser-induced demagnetization, (ii) the all-optical switching
of the magnetization with fs laser pulses, and (iii) the ability to optically generate
and exploit fs spin currents in nanostructured magnetic systems.

Conceptual Introduction

The 1990s saw an entirely new development, facilitated by the commercial avail-
ability of fs pulsed laser sources, which began to be used to explore the ultimate
limits of magnetization dynamics. A major outcome of these studies is that we no
longer see optics as merely a powerful tool to read the magnetic state but also as
a very efficient tool to manipulate it. As such, a new field of opto-magnetism has
emerged, in which magnetic ordering is affected by the light.
Opto-magnetism provides a method to control the magnetic state beyond the
traditional precessional dynamics as described most conveniently by the Landau-
Lifshitz-Gilbert equation of motion [48]

dM α dM
= −γ μ0 (M × Heff ) + M× . (35)
dt M dt

Here, an effective field H eff is introduced that applies a precessional torque on the
magnetization M as described by the first term on the right-hand side, where γ is
10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films 503

the (positive) gyromagnetic ratio and μ0 the vacuum permeability. The second term
describes a phenomenological damping via the Gilbert damping parameter α, which
allows the system to dissipate energy and thereby converge to the lowest energy
state, in which M is aligned with H eff . The precessional dynamics in ferromagnets
proceed at frequencies of up to 10s of GHz for applied magnetic fields not exceeding
1 MA/m (μ0 H ∼ 1 Tesla) and conserves the magnitude of the local magnetization.
Opto-magnetism provides a route to change not only the orientation but also the
magnitude of M, and it does so on a time scale that is typically up to three orders of
magnitude faster.
A first (basic) mechanism to exploit pulsed lasers to manipulate ferromagnetic
order is laser heating. Pulsed lasers have enough energy per pulse to heat up a
ferromagnetic thin film of any material and with a thickness comparable to the skin
depth of the light, to well above its Curie temperature. Increasing the temperature
will lower the equilibrium magnetization, and the system will strive to re-establish
thermal equilibrium between the spin- and other degrees of freedom, lowering its
magnetization. Finding the relevant time scale associated with losing magnetic
order was one of the fundamental quests that inspired researchers in the 1990s.
Beaurepaire and co-workers were the first to find that the relevant time scale was
well below a picosecond [4]. This rapid loss of magnetization was referred to as
ultrafast demagnetization.
Effects of laser heating can be even more spectacular if they provoke a transition
between contrasting magnetic phases. A trivial example is laser heating through the
Curie temperature, driving a system from the ferromagnetic to the paramagnetic
state. Then, the magnetization can regrow in the opposite direction upon cooling
down when a static reversed magnetic field is applied [49], as is shown in Fig. 11a.
A more interesting situation occurs for materials like FeRh, which can be grown

Fig. 11 (a) Recovery of the magnetization of GdFeCo alloy after a full laser-induced demagneti-
zation in the presence of a applied magnetic field in the original direction (triangles up), zero field
(squares), and reverse field (triangles down). (Figure taken from Ref. [49]. with kind permission
of APS) (b) Laser-induced growth of magnetization in FeRh as measured by TR-MOKE (filled
symbols), compared to the transient reflectivity ΔR. (Figure taken from Ref. [50] with kind
permission from APS)
504 M. L. M. Lalieu and B. Koopmans

such that it is antiferromagnetic around room temperature but ferromagnetic at


higher temperatures. In this case, fs laser heating could be expected to cause a laser-
induced growth of ferromagnetic order, as is indeed observed [50, 51]; see Fig. 11b.
Consensus has been reached that the final magnetic phase transition is completed
within several tens of ps [52, 53].
Apart from affecting the equilibrium magnetization, fs laser heating can be used
to optically generate magnetization precession [54,55,45], as was shown in Fig. 10,
or to launch standing spin waves [56]. In these approaches, the laser-pulse excitation
causes a rapid sub-ps change of the magnetic anisotropy, after which the dynamics
is governed by the LLG equation, Eq. (35). A similar approach has been used to
optically modify the interface exchange bias between a ferromagnetic film and an
adjacent antiferromagnetic film [57].
Besides the thermal effects, the electromagnetic photon field associated with the
laser pulse itself can induce changes in the magnetic state, as mediated by spin-
orbit coupling [58, 59]. Although such effects certainly exist, their relevance to
explaining the experimentally observed sub-ps dynamics in metallic systems is not
always evident, as will be discussed in section “Theories for Femtosecond Demag-
netization”. However, in special materials spectacular effects have unambiguously
been identified. For instance, in 2017 it was shown that linearly polarized fs laser
pulses break the degeneracy between meta-stable magnetic states in transparent
nonmetallic iron garnets, leading to optically induced switching with exceptionally
low energy absorption [60].
All-optical switching of magnetization (AOS) in metallic systems was discovered
by Stanciu et al. in 2007 [5]. It was shown that magnetization of GdFeCo alloys can
be “written” by a single fs laser pulse in a helicity-dependent (HD) manner, i.e.,
the final state (magnetization up or down) is determined entirely by the helicity of
the light (left- and right-handed circular polarization), as was illustrated in Fig. 1b.
Later, it was found that helicity-independent single-pulse toggle switching played
a decisive role in these seminal experiments, governed by a dynamic exchange
mechanism. In this scenario, strong exchange forces (corresponding to a field
up to > 1000 T) drive an ultrafast reversal of the oppositely oriented magnetic
moments at the Gd and Fe/Co atomic sites [6] (see section “All-Optical Switching
of Magnetization”).
From a conceptual point of view, it is of relevance to introduce the inverse
Faraday effect (IFE), which is another effect of the photon field. It elegantly
reflects the symmetry between magneto-optics and opto-magnetism. The IFE is
reciprocal to the ordinary Faraday effect and was proposed in the 1960s. It was
first experimentally demonstrated for a nonabsorbing material by van der Ziel et al.
[61]. In the language of Eq. (13), the IFE describes the generation of a temporary
(quasi-static) magnetization M in the material according to

M = αE × E ∗ , (36)

where E is the amplitude of the oscillating (optical) electrical field as defined before,
E ∗ its complex conjugate, and α a material constant characterizing the size of
the IFE in the material. It should be emphasized that the cross product E × E ∗
10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films 505

vanishes for linearly polarized light. For circularly polarized light, we see that the
induced magnetization reverses upon reversal of the helicity. Furthermore, it should
be stressed that the induced M is coherent in the sense that it is setup by the laser-
field and decays within a decoherence time. It has been suggested that the IFE
may be of relevance for HD-AOS of metallic ferrimagnets, which is discussed in
section “All-Optical Switching of Magnetization”.
Laser excitation has been demonstrated to be a source of laser-induced spin
currents – whether by spin-polarized ballistic electrons, of “superdiffusive”
nature, or driven by the demagnetization or thermal gradients. These spin
currents trigger a variety of magnetization dynamics, further elaborated on in
section “Laser-Pulse-Excited Spin Currents”.

Ultrafast Laser-Induced Loss of Magnetic Order

We now focus on the process of fs laser-induced demagnetization. A historical


review of early experiments is followed by a collection of key-experimental results
that should guide the explanation. Then, the issue of conservation of angular
momentum is revisited, after which theories proposed to explain the ultrafast
demagnetization are analyzed. Finally, a selection of more recent advances in the
field is discussed.

Experimental Demonstration of Laser-Induced Demagnetization


Experiments addressing the time scale at which magnetic order can be quenched
after sudden (laser) heating started in the 1980s [62]. The first real time-resolved
experiments were performed in the early 1990s by Vaterlaus and co-workers [63,
64]. Ferromagnetic gadolinium, a rare-earth ferromagnet with a Curie temperature
just below room temperature, was heated by 10 ns pump pulses. The magnetic order
was measured by analyzing the spin-polarization of the electrons photoemitted by
60 ps probe pulses. The relaxation time of the magnetization was found to be
100 ± 80 ps. A similar spin-lattice relaxation time was calculated by Hübner and
Bennemann [65].
This was the background for the pioneering experiments by Eric Beaurepaire,
Jean-Yves Bigot, and co-workers in 1996. They were the first to study demagneti-
zation by using <100 fs laser pulses to address the strongly nonequilibrium regime.
Applying a pump-probe TR-MOKE approach to nickel thin films, they observed
a quenching of the magneto-optical contrast by approximately 50% within a few
hundred fs (Fig. 1a). This demagnetization time scale was orders of magnitude faster
than any earlier experimental estimate or prediction. The result was confirmed soon
after by fs MSHG studies (section “MO Spectroscopy”) by Hohlfeld et al. [47], as
shown in Fig. 12a, and by spin-resolved two-photon photoemission experiments by
Scholl et al. [66]. These novel findings triggered enormous interest and gave birth
to the field of femtomagnetism.
Beaurepaire et al. interpreted their results phenomenologically in terms of the
three-temperature model (3TM) sketched in Fig. 12b, c. In this model, the material is
described in terms of three interacting thermalized reservoirs, provided by electrons,
506 M. L. M. Lalieu and B. Koopmans

a b

gep
Te Tp
Ce Cp
ges gsp
Ts

Cs
Fig. 12 (a) Induced MSHG signal as a function of pump-probe delay time at different laser
fluences measured for nickel. Displayed fluences normalized to 1.00 ≈ 6 mJ/cm2 . (Figure taken
from Ref. [47]. with kind permission from APS) (b) Schematic diagram of the evolution of Te (t)
(red), Tp (t) (blue), and Ts (t) (green), as well as the corresponding M(t)/Msat (green, top panel),
after pulsed laser excitation at delay 0, calculated using the 3TM (Eqs. 41 and 42), and setting
gsp = 0. Demagnetization time τM and e-p equilibration time τE are indicated. Note that a peak
in Ts (t) corresponds to a dip in M(t). (c) The 3TM, showing the three interacting reservoirs
with temperature Ti and heat capacity Ci and that exchange heat via a coupling constant gi,j
(i, j = e, p, s), see discussion Eq. (42)

their spins, and the lattice (phonons). To each of them, a temperature is assigned
(discussed in more detail later). In terms of this model, the laser pulse causes
an ultrafast rise of the electron temperature, which subsequently cools down by
exchanging heat with the lattice as mediated by electron-phonon coupling. The
electron-phonon equilibration time is denoted by τE (see Fig. 12b) and is typically
of the order of ∼0.5 − 1.0 ps. In parallel, the spin system is heated up by interaction
with the electrons and the lattice, causing the rapid loss of magnetization with a
demagnetization time τM , and finally equilibrates with both of them. The surprising
outcome was that the spin temperature follows the electron temperature faster
than the lattice does. It hereby reaches a peak – corresponding to a dip in the
magnetization – after approximately half a picosecond. Successively, the spins cool
10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films 507

down – corresponding to recovery of the magnetic order – to thermalize with both


electron and lattice system (see Fig. 12b).
Although different experiments had independently demonstrated an ultrafast loss
in magnetic contrast, the question arose as to how well these signals represented the
evolution of the magnetization while still in the strongly nonequilibrium state. This
question can be rephrased as the relative importance of the term proportional to
ΔF (t) in Eq. (30). Koopmans et al. triggered this discussion by showing that the
MO rotation and ellipticity displayed a different behavior for epitaxial Ni thin films
on Cu during the first 100s of fs in a TR-MOKE experiment, while at longer time
scales, they fully converged [67]; see Fig. 13a. Via the Kramers-Kronig relation,
the difference in rotation and ellipticity is reflecting pump-induced changes in
spectral MO features. These differences were attributed to state-filling effects, also
referred to as “dichroic bleaching,” arguing that even if optical excitations are spin-
conserving, the magneto-optical spectrum should change irrespective of whether
the magnetic moment would be conserved [68]. Similar state-filling effects were
found in experiments by Regensburger et al. [69] and Guidoni et al. [70], as
shown in Fig. 13b. Theoretical estimates of the effects right after laser excitation
by Oppeneeer et al. matched the experimental observation in Ni rather well [71].
The effects that contribute to the ΔF (t) term have often been referred to as
“optical artifacts.” However, it should be emphasized that even though they do
not reflect the dynamics of the net magnetization, they may carry highly relevant
information about the transfer of spin angular momentum. As an example, in alloys,
compounds, or thin film multilayers, femtosecond optical excitation can induce
transfer of spin angular momentum between different magnetic sublattices, an effect
that has been dubbed OISTR (optically induced intersite spin transfer) [72, 73].
Despite the fact that it is commonly accepted that care has to be taken
when interpreting time-resolved studies of magnetization dynamics in the strongly

Fig. 13 (a) Normalized rotation (filled diamonds) and ellipticity (open circles) channels measured
by TR-MOKE for a Ni thin film, showing clear differences during the first few hundred fs. (Figure
taken from Ref. [67] with kind permission from APS). (b) Similar results on CoPt3 ; rotation (open
diamonds) and ellipticity (filled circles). (Figure taken from Ref. [70] with kind permission from
APS). (c) Calculated dichroic bleaching for nickel at a similar excitation density fexc (defined as
the number of optically excited electrons per atom) as used in the measurements in (a). The vertical
dashed lines indicate the photon energy used in the experiment. (Figure taken from Ref. [71])
508 M. L. M. Lalieu and B. Koopmans

nonequilibrium regime, it has become clear that a genuine sub-ps demagnetization


does occur. The artifacts have been identified to be of relevance only in specific
cases, such as during pump-probe overlap, or in cases where an observed tensor
element is close to a zero crossing with respect to its wavelength dependence. As
an illustrative example, one could perform a TR-MOKE experiment at a probing
wavelength chosen such that the MO spectrum displays a zero crossing, e.g., 1.8 eV
for Δθ in Fig. 13c. In such a case, any change of the MO spectrum, for example, due
to repopulating electronic states or a temperature dependence of the electronic band
structure, would imply a finite MO signal after laser excitation and thus an infinite
relative change that is not directly reflecting a change in magnetic moment.

Key Observations in Laser-Induced fs Demagnetization


In the following, we present the cumulative evidence for sub-ps magnetization
dynamics that emerged since 2000, as well as key observations that have guided
further advances in microscopic understanding. Additional to the elementary fer-
romagnetic 3d transition metals, sub-ps laser-induced demagnetization has been
observed for many different materials: alloys such as permalloy and CoPt3 [74], the
ferromagnetic rare-earth materials Gd and Tb [75], and the ferrimagnetic transition-
metal rare-earth alloys such as GdFeCo that also display AOS (section “All-Optical
Switching of Magnetization”). Furthermore, ultrafast demagnetization is observed
for several ferromagnetic oxides [76, 77], Heusler alloys [76, 78], and magnetic
semiconductors such as InMnAs [79,80], and several mangantites and other oxides,
e.g., Refs. [81, 77].
In addition to studies that show the ultrafast quenching of the magnetization, Rhie
et al. reported on a collapse of the exchange splitting in Ni within 300 fs as measured
by time-resolved photoemission spectroscopy [82]. A true breakthrough has been
established by pushing XMCD (section “MO Spectroscopy”) to the fs regime by
an approach called femto-slicing [83]. Dürr et al. showed that the magnetization
dynamics at a sub-ps time scale measured using XMCD was very similar to the one
probed by magneto-optics. It was demonstrated that the spin and orbital momenta
evolve in quite a similar fashion, although subtle differences were observed [84].
A key result is depicted in Fig. 14a. Table-top extensions of fs XMCD have been
used to measure ultrafast demagnetization for different ferromagnetic thin films,
again resulting in a demagnetization time of typically 100–300 fs [34]. Its element
specificity has been used to probe subtle differences in demagnetization rates and
interatomic spin transfer, between Ni and Fe in NiFe alloys [33].
Complementary evidence for genuine sub-ps demagnetization came from the
observation of THz emission. A sudden change of magnetic dipole moment at
a ps time scale should act as an antenna for electromagnetic radiation at a THz
(= ps−1 ) frequency. This can be measured in the far field by free-space electro-
optic sampling, as first shown by Beaurepaire in 2004 for Ni films [87]. Work by
other groups reported similar observations [88]. Simultaneous detection of the THz
emission and transient MO response for Co, GdFCo, and NdFeCo revealed a very
similar time evolution of the magnetic signal [89].
Information on mechanisms underlying the sub-ps demagnetization has been
obtained by comparing different excitation methods. In this respect, the demon-
10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films 509

(a)
0.8
(b)
Spin and orbital moments ( ħ per atom)

Sz

0.6 th = 280 ± 20 fs

0.4 –55%
Lz
2.5 mJ/cm2
3.0 mJ/cm2
0.2 = 220 ± 20 fs 3.5 mJ/cm2
th
4.0 mJ/cm2
4.5 mJ/cm2
–67% 5.0 mJ/cm2

0.0
0 1 2
Dela y (ps)

(c)

Fig. 14 (a) Spin (Sz , red) and orbital (Lz , blue) magnetic moments as a function of the delay time
between the laser pump and the X-ray probe pulses, extracted using sum rules. The measurement
is performed on a ferromagnetic Co0.5 Pd0.5 thin film using the Co L2,3 absorption edges. (Figure
taken from Ref. [84] with kind permission from APS). (b) Fluence dependence of the laser-induced
sub-ps demagnetization of a 15 nm Ni film at room temperature. Figure taken from Ref. [85]. (c)
Time-resolved magnetization following single-cycle THz pumping of a 5 nm CoFeB film. Inset:
CoFeB demagnetization amplitude as a function of the THz peak field. (Figure taken from [86])

stration of ultrafast demagnetization by pulses of hot electrons is of relevance


[90, 91]. In a multilayered structure with a Pt absorption top layer, separated by
a Cu transport layer from a ferromagnetic [Co/Pt]N bottom layer, bunches of
hot electrons were created by fs laser pulses, which ballistically traversed the Cu
spacer, before depositing their kinetic energy in the ferromagnetic layer [90].
A delayed demagnetization proceeding at almost the original rate was observed,
and an electron velocity close to the Fermi velocity in Cu could be deduced.
Most importantly, this rules out the laser optical field as an essential ingredient
for the ultrafast demagnetization, because the demagnetization starts well after
the optical pump pulse is over. In another experiment, it was demonstrated that
ultrafast demagnetization is possible by electronic Joule heating, induced by a
510 M. L. M. Lalieu and B. Koopmans

ps electrical current pulse [92]. These experiments are indicative that neither the
photon field nor hot electrons with eV energies are an essential ingredient for sub-
ps demagnetization. A similar conclusion can be drawn from a TR-MOKE study by
Roth et al. on a Ni thin film at different pump laser fluences [85], shown in Fig. 14b.
The demagnetization slows down at increasing laser fluence. At the highest fluence,
a rather continuous demagnetization takes place up to a ps, well after the driving
laser field has decayed and electrons have thermalized (both typically 50–100 fs).
Yet another way of exciting sub-ps demagnetization is by using single-cycle THz
pulses. Bonetti and co-workers performed THz-pump MOKE-probe experiments on
crystalline Fe and amorphous CoFeB [86]. Although the amount of demagnetization
was very small, typically 0.1%, the drop in magnetization followed quite accurately
the dissipated power as expressed by the THz driven current J (t) and the THz
electric field E(t); see Fig. 14c.
Although the foregoing results seem to favor a local and thermodynamic picture,
there is growing evidence that this is not the whole story. The community has
become aware of the role of optically induced spin transport, first demonstrated by
Malinowski et al. [10]. There is growing evidence that the demagnetization process
proceeds faster for ultrathin films on top of a metallic substrate as compared to
insulting substrates, due to spin angular momentum being efficiently transported
out of the ferromagnetic layer [46]. Yet, even when confining the metallic system in
a nonconductive environment to prevent spin transport, ultrafast demagnetization at
a 100 fs timescale is consistently being seen [93].

Conservation of Angular Momentum Revisited


Conservation of angular momentum plays a key role in laser-induced demagne-
tization. Let us consider a closed (ferromagnetic) system of finite dimensions,
interacting with a laser pulse. Applying a magnetic field H at an angle to the
magnetization M results in a torque τ , which causes precessional dynamics
according to the LLG equation (Eq. 35). Consequently, the magnetization M and the
associated angular momentum J precess together. On the other hand, if no external
torque were applied, J will be conserved, i.e., dJ /dt = τ = 0. In that case, we can
write

J = Le + S e + Llat + J EM = constant, (37)

where we separated the electron contribution into an orbital and a spin part, Le and
S e , Llat is the angular momentum provided by the lattice, and J EM is the angular
momentum in the photon field.
A fundamental bottleneck in ultrafast changes of the magnetization is provided
by the need to conserve the total J while changing the magnetic moment:

μ = μB (Le + gS e ). (38)

For the materials considered here, g ≈ 2. In the ground state, the magnetic moment
is strongly dominated by Se , since the orbital momentum is significantly quenched
10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films 511

by crystal-field effects. An obvious channel for transfer of angular momentum


would be exchange of spin- and orbital momentum mediated by spin-orbit coupling.
However, it can easily be verified that this can only lead to quenching half of the
magnetic moment. Moreover, fs-XMCD has shown that both Le and S e decay in
a similar fashion, rather than angular momentum being transferred from S e to Le
[83, 84]; see Fig. 14a.
A second channel would be spin-lattice relaxation. This is a very familiar process,
related to the famous Einstein-de Haas effect. In this experiment, a demagnetized
ferromagnetic bar, i.e., a multidomain state with net zero magnetic moment, is
suspended by a thin rope and is free to rotate. Upon applying a magnetic field
along the vertical axis, the bar starts to rotate due the angular momentum transfer
between the electron system and the lattice [94]. Einstein and de Haas used this to
estimate the g-factor, by applying Eqs. (37) and (38) [95]. Note that in the absence
of spin-lattice relaxation, i.e., without (Gilbert) damping, α = 0, the local magnetic
moments will keep on precessing in the local effective field while retaining their
net angular moment, so that the bar would not start to rotate. Despite this analogy,
however, this channel was originally considered too slow to account for sub-ps
demagnetization.
A third channel is direct exchange of angular momentum with the photon
field. However, it was argued already in 2000 that under experimental conditions,
there would be too few photons to cause the observed change in magnetization
[67]. Moreover, from experiments discussed before, we concluded a non-decisive
role of the photon field. In fact, this channel can obviously not account for
any demagnetization when considering a (quite common) situation of having the
magnetization perpendicular to the propagation direction of the light, which sets the
direction of the light’s angular momentum.
Having rejected all three channels, there might be something wrong in our
naive interpretation. At least part of the answer is hidden in the notion that the
Hamiltonians that we generally use to describe the spin scattering processes do not
conserve angular momentum! Typically, we describe infinite systems with periodic
boundary conditions and cubic or other nonspherical arrangements, in which J is
not a good quantum number.2 If, in contrast, we would have set up a Hamiltonian
describing all degrees of freedom for an isolated finite crystal, including the global
rotation of the crystal as a whole, the system would be invariant under its overall
rotation, such that the angular momentum would be conserved. Evaluating the
associated laser-induced dynamics is a formidable task which to the best of our
knowledge has not been reported, although the consequences have been referred to
in Refs. [96, 97].
To explain the consequences in a didactic way, let us consider an atom with
d-orbitals and a square arrangement of nearest neighbors. We restrict ourselves to
the dxy and dx 2 −y 2 states, which become nondegenerate due to the crystal field with
a square symmetry, with an energy difference ΔE. We enforce a single electron√in
an initial state given by a linear combination of atomic orbitals, (dxy + idx 2 −y 2 )/ 2
with magnetic quantum number m = 2. This state is not √ a stationary eigenstate. Its
time evolution corresponds to (dxy +i eitΔE/h̄ dx 2 −y 2 )/ 2, which displays a periodic
512 M. L. M. Lalieu and B. Koopmans

beating between the m = 2 and m = −2 state. The apparent conclusion is that the
angular momentum is not conserved! This is due to the torque applied by the crystal
field on the d-orbitals, while we neglected in our Hamiltonian the back action on
the neighboring nuclei. If we would have included the motion of the nuclei in our
Hamiltonian, the total angular momentum would have been exactly conserved by
compensating contributions from the electron and the nuclei.
Extending this picture to laser-induced demagnetization requires many additional
interactions to be included in our Hamiltonian, including spin-orbit coupling as an
essential ingredient. Upon laser heating, the original angular momentum associated
with the magnetic ordering is transferred to a local twist of the crystal within an
area comparable to the size of the laser spot. This local twist will propagate radially
outward as a torsional shockwave. Ultimately, this leads to a final (though minute)
rotation of the crystal, c.q. its holder, cq., the optical table, etc., in a fashion similar
to a picture sketched by Fähnle et al. [98].
With the notion in mind that our Hamiltonian generally does not necessarily
conserves angular momentum, we analyze the possible underlying elementary
scattering processes. A relevant one is the Elliott-Yafet spin-flip scattering, in which
electrons have a finite chance to flip their spin upon scattering at impurity sites or by
emission or absorption of a phonon (in a e-p scattering process) [99]. The resulting
spin-flip probability is a consequence of the fact that spin (nor J ) is not a good
quantum number in our crystals. Thereby, eigenstates of our Hamiltonian are of a
mixed spin character of the type

ψk,σ = ak,σ |k, ↑> +bk,σ |k, ↓>, (39)

for σ =↑, ↓ and where labeling is chosen such that for all states b < 0.5. Then, a
spin-mixing parameter < b2 > can be calculated after weighing it over all possible
scattering processes. This spin-mixing parameter scales with Z 4 , where Z is the
ionic charge in units of e, and typically b2 << 1 for materials under consideration.
If an electron scatters from a state k to k  , the spin can generally not be conserved,
which finally leads to a spin-flip probability asf of the order of the bk2 after a weighted
averaged over the Brillouin zone. We emphasize that it is an incorrect picture to state
that (all) spin angular momentum associated with the electronic spin-flip process is
transferred to the specific phonon emitted or absorbed.
Interestingly, this process of EY type of scattering with phonons can be trivially
extended to the process of electron-electron (e-e) scattering. Also, in this case, the
process can be quantified by a finite spin-flip probability during the e-e scattering
event – an approach adapted e.g. by Steil et al. [97] and further elaborated on in
the next section. Similar to the phonon-mediated Elliott-Yafet process, the angular
momentum will be dumped in the rotational degrees of freedom of the crystal.

Theories for Femtosecond Demagnetization


Already from the 1980s, electron and lattice dynamics in metals after fs or ps
laser excitation has been investigated. Laser photons create electron-hole pairs, with
electron energies E up to the photon energy of a few eV above the Fermi level, EF .
The inelastic lifetime τ0 of these electrons is described by
10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films 513

τ0 ∝ (E − EF )−2 . (40)

Typically, life times are just a few fs for energies of more than an eV above EF .
Electrons kick out other electrons from the Fermi sea via Coulomb scattering,
starting a cascade of increasingly more excited electrons while conserving the total
excess energy. After a typical thermalization time of τT = 100 fs, the occupation
of electronic states resembles a Fermi-Dirac distribution again, a process that can
be followed by time-resolved photoemission [82,100]. For ferromagnetic materials,
the life time for majority and minority carriers can be significantly different due to
the different phase space for spin-conserving Coulomb scattering, as first measured
by Aeschlimann et al. by spin-dependent two-photon photoemission in 1997 [101].
This spin-dependent lifetime is of crucial relevance for superdiffusive spin transport,
discussed in section “Laser-Pulse-Excited Spin Currents”.
After thermalization of the electron system, a two-temperature model can be
used to describe the dynamics of the heated electron system in contact with the
cooler lattice. Thermal reservoirs are introduced for the electrons and the lattice
(phonons), as in the 3TM sketched in Fig. 12c. Each bath is assumed to be in
full internal equilibrium, and a temperature, Te and Tp , respectively, is assigned
to it. Furthermore, both baths are assigned a heat capacity, Ce and Cp , and a
coupling constant gep is introduced to account for heat exchange, microscopically
mediated by electron-phonon (e-p) scattering processes. Thus, a simple set of
coupled differential equations is derived, including a source term that describes the
heating of the electrons by the laser pulse. The solution describes a process in which
initially the electrons are heated, after which electrons and lattice equilibrate at a
time scale τE ∼ 0.5 ps for the ferromagnetic transition metals.
For a phenomenological description of laser-induced demagnetization, Beaure-
paire et al. used an extension of the two-temperature model, which has been dubbed
the three-temperature model (3TM) [4]. Like the electron and lattice system, the
spin system is assumed to be in full thermal equilibrium and is assigned a “spin
temperature.” Although not a trivial concept, it can most easily be defined by
relating the instantaneous magnetization M(t) at certain time t to the equilibrium
magnetization at temperature at T:

Meq. (Ts (t)) = M(t). (41)

The resulting three coupled differential equations read

dTe
Ce = gep (Tp − Te ) + ges (Ts − Te ) + P (t),
dt
dTp
Cp = gep (Te − Tp ) + gsp (Ts − Tp ),
dt
dTs
Cs = ges (Te − Ts ) + gsp (Tp − Ts ), (42)
dt
514 M. L. M. Lalieu and B. Koopmans

where Cs is the spin specific heat, diverging at TC , ges and gsp are the electron-
spin and spin-lattice coupling constants, respectively, and P (t) is the source term.
In the limit of low laser fluence and zero pulse length, i.e., instantaneous heating
of the electron gas, and neglecting the spin specific heat, an analytical expression
for the time evolution of the magnetization ΔM(t) has been derived in terms of τE
and τM [102, 45]. This expression can be conveniently used to fit and parameterize
experimental data sets [103]. Despite its use for phenomenological understanding
and as a quantitative fitting tool, the 3TM does not provide any microscopic insight.
Next, we will discuss theoretical approaches that have tried to pinpoint the key
microscopic mechanism of sub-ps demagnetization.
As briefly introduced before, it has been suggested that the photon field might
play a crucial role. A first mechanism is a direct effect, in which the magnetic
moment in the excited state after optical absorption of a photon differs from
the one in the ground state. However, effects will be relatively small in the
prototype transition-metal (TM) ferromagnets since the orbital magnetic moments
are quenched. Moreover, it has been argued that the number of photons in
typical experiments is far too small to account for the observed magnetization
dynamics [67].
In 2000, it was argued that a laser field in combination with electron correlation
and spin-orbit coupling could lead to an ultrafast demagnetization at tens of fs
[104, 105, 106]. Later, Bigot et al. conjectured that the material polarization induced
by the photon field may interact coherently with the spins, corresponding to a
mechanism that has its origin in relativistic electrodynamics [59]. In passing, it
should be mentioned that the latter results have been criticized in Ref. [107].
Moreover, the experimental effects reported in [59] were only seen during the
presence of the laser pulse and left no imprint on the magnetization afterward.
As mentioned in the previous section, spin-orbit coupling leads to spin-mixing in
the electronic eigen states, which is a source of spin-flip scattering when electrons
scatter with other quasi particles. Koopmans and co-workers have proposed a
theoretical framework to incorporate Elliott-Yafet type of spin-flip scattering with
phonons [55, 103] or impurities [55] into the earlier mentioned 3TM. Within this
microscopic three-temperature model (M3TM), a simple model Hamiltonian is used
to derive three coupled differential equations for the time evolution of Te , Tp , and the
normalized magnetization m = M/Ms (directly related to Ts according to Eq. (41)).
The model is schematically depicted in Fig. 15a.
In the M3TM, the electron system is described by spinless free electrons, with
a constant density of states near the Fermi level (DF ), leading to Ce = γ Te with
the constant γ ∝ DF . Phonons are described within the Einstein model, or mapped
on a Debye model, in both cases described by the Debye (phonon) energy ED ,
and the phonon number of modes per atom Np . The phonon heat capacity Cp
is assumed to be independent of T , a reasonable approximation near and above
the Debye temperature. Finally, spin excitations are treated using a spin S = 1/2
mean-field Weiss model, introducing the atomic magnetic moment μat and the Curie
temperature TC , directly related to the mean-field exchange parameter J . The spin
specific heat is neglected, Cs = 0, so that the evolution of Te and Tp does not
10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films 515

(a) (b)
E (½ + n) h p

+½ e
EF kTe
ex
-½ asf p
DOS e
Electrons Spins Phonons

Fig. 15 Schematic diagram of the M3TM. (a) Representation of the electron, phonon, and spin
system, including a scattering event in which an electron loses kinetic energy, a phonon is emitted,
and a spin is flipped to reduce the magnetization, while conserving energy. (b) Elliott-Yafet spin-
flip mechanism, in which a spin is flipped during an e-p scattering event

depend on Ts . In the interaction Hamiltonian an additional parameter is avoided


by assuming infinitely efficient e-e Coulomb scattering, leading to instantaneous
electron thermalization. To account for e-p scattering, a coupling constant λep is
introduced. Note that e-p coupling is most efficient for zone-edge phonons, which
validates the use of a single-phonon energy with a magnitude of approximately ED .
Finally, and most importantly, spin-flip electronic scattering upon phonon emission
or absorption is assigned a prefactor asf λep , where asf is the spin-flip probability,
see Fig. 15b.
Applying Fermi’s Golden rule to the scattering processes and implementing the
outcome in a set of Boltzmann equations, a set of coupled differential equations is
derived. The first two, for Te and Tp , are identical to the ordinary 3TM (Eq. (42),
with the second terms on the right-hand side set to 0). The third equation, describing
the magnetization dynamics, reads


dm Tp mTC
= Rm 1 − m coth , (43)
dt TC Te

Thus, a compact equation for the evolution of the magnetization dynamics m(t) is
found, in which the rate of demagnetization is described by the parameter

8asf gep kB TC2 Vat


R= 2
, (44)
μat ED

(in units of s−1 ), where Vat is the atomic volume, and gep ∝ λ2ep can be expressed in
elementary model parameters (see Supplementary Information of Ref. [103] for
details). The parameter asf is proportional to < b2 > in Eq. (39), highlighting
the importance of spin-orbit interaction. One readily sees that no magnetization
dynamics occurs, i.e., dm/dt = 0, if asf = 0, as expected.
Two other approaches, although rather contrasting at first sight, reproduce
magnetization dynamics very similar to the M3TM. Firstly, the Landau-Lifshitz
Bloch (LLB) equations provide an extension of the LLG equation, by adding
516 M. L. M. Lalieu and B. Koopmans

an additional differential equation for the longitudinal component of M. Atxitia


et al. treated the longitudinal magnetization dynamics after pulsed laser heating for
arbitrary S = n/2 (where n is an integer). They derived an analytic differential
equation for the longitudinal magnetization relaxation for S = 1/2 and showed that
it can be exactly mapped on Eq. (43) [108]. As such, the M3TM can be considered
a microscopic quantum-mechanical derivation of the longitudinal part of the LLB.
Consequently, identical dynamics is reproduced.
Secondly, atomistic LLG (ALLG) models produce very similar M(t) traces as
well [109]. In ALLG calculations, a large grid of atomic magnetic moments of fixed
magnitude is considered, in which neighbors couple via interatomic exchange, and
stochastic thermal noise is introduced to account for (electronic) temperature effects.
Angular momentum dissipation to the lattice is described by the phenomenological
Gilbert damping parameter α, Eq. (35). Assuming that α at the atomic scale has the
same magnitude as the one describing mesoscopic magnetization precession, very
similar magnetization dynamics upon laser heating is obtained compared to LLB
and M3TM [109]. The ALLG approach lacks both the numerical simplicity and the
quantum mechanical footing of the M3TM but, on the other hand, allows for treating
spin fluctuations, stochasticity, and the inclusion of realistic spin wave modes [110].
The assumption that the microscopic and mesoscopic α are identical may seem a
bold statement, but it has interesting consequences. In ALLG, the demagnetization
time scales as α −1 , and as such α plays a role similar to asf in the M3TM. Actually,
also in the M3TM, a simple relation between the mesoscopic α and τM has been
derived, by a simplified quantum mechanical calculation of damping in precessional
dynamics based on the same Hamiltonian used in the M3TM. The result thereof
is [55]

h̄ 1
τM ≈ c0 , (45)
k B TC α

where c0 is a constant of the order of unity, typically ∼1/8, but depending on


details. Taking a Ni thin film as an example, using α equal to a few hundredths and
TC = 630 K, one readily obtains τM ∼ 100 fs in close agreement with experimental
observations. An intuitive interpretation of Eq. (45) is that atomic fluctuations in
the local magnetization damp out (or thermalize) at the same rate as a macroscopic
magnetization would do in a field set by the atomic exchange field (governed by TC
and typically 1000s of Tesla) and governed by the macroscopic α.
Theoretical predictions for Elliott-Yafet spin-flip probabilities of transition-metal
ferromagnets have been obtained by ab initio calculations of the Eliashberg function
by Carva et al., while order of magnitude estimates were obtained by calculating the
< b2 > parameter by Fähnle et al., [111]. Resulting values are typically 0.05–
0.10, in close agreement with the interpretation of experiments using the M3TM
[85] (section “Towards Quantitative Understanding”). Nevertheless, the relevance
of this agreement has been disputed. In a rigid band-structure calculation adapting
a Stoner picture of magnetism (in which magnetization is changed by variation
of the atomic magnetic moment), Carva et al. find a negligible demagnetization
10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films 517

using the calculated ab initio value, whereas Schellekens et al. argue that close
agreement with experimental results is obtained by implementing them in the Weiss
or Heisenberg picture (in which the magnetization dynamics is due to thermally
disordered orientation of fixed atomic moments) [112].
The basic M3TM makes use of severe approximations that can be lifted when
desired. One can implement a more realistic electronic density of states, phonon
dispersion, or extensions from S = 1/2 to arbitrary spin multiplets. Also, the
description of itinerant magnetism in a pure Weiss picture may be considered
inadequate. The s-d model applied to ultrafast magnetization dynamics in Ref. [113]
provides an interesting step beyond.
We treated phonon-mediated Elliott-Yafet spin-flip scattering in quite some
detail. Spin-flip e-e scattering is another possibly relevant scattering mechanism, in
which one of the two mutually scattering electrons flips its spin. Although it could be
argued that this process does not conserve angular moment, we highlighted before
that it will be facilitated by the non-spin-conserving character of the Hamiltonian
and will lead to a dump of angular momentum in the rotational degrees of motion of
the crystal. Steil et al. included this type of Coulomb scattering, without the need of
referring to a phononic spin bath [97, 114]. They implemented this in a Stoner-
like description in which electrons are scattering between different spin bands.
Reasonable agreement with experiments was found assuming spin-flip probabilities
of 0.15 and 0.30, for Co and Ni. From this, a μT -model emerged, in which
dynamics is described in terms of equilibration of temperatures and chemical
potentials of the electron subsystems simultaneously [115]. This approach bares
some similarities with the s-d model of Ref. [113].
Finally, the increase of computational power and refinement of codes have fueled
activities on many-body theory and ab initio calculations. Ultimately, this would
enable to trace the angular momentum transfer from a microscopic perspective.
Töws and Pastor performed small-cluster calculations distinguishing local and
itinerant electrons and exploring the interplay between spin-orbit interactions and
interatomic hopping, as well as the role of the photon field [96]. Realistic time
scales of demagnetization of typically 100 fs were obtained. Although the lattice
degree of freedom was not included in this work, the authors stated that the leaking
angular momentum would manifest itself as a global rotation of the rigid lattice if it
were explicitly included in the calculation.
Fully ab initio calculations of laser-induced magnetization dynamics have been
reported by Sharma and co-workers [116, 117]. They employed time-dependent
density functional theory in an all-electron solid-state code for the case of uncon-
strained noncollinear spins, as driven by short pulses of electro-magnetic radiation.
Despite the fact that the amount of demagnetization observed was at least an order
of magnitude smaller than realistic experiments would show and the phonons were
not included yet, this field is rapidly making progress.

Towards Quantitative Understanding


We conclude our treatment of laser-induced fs demagnetization by a selection of
studies that aimed at a more quantitative understanding.
518 M. L. M. Lalieu and B. Koopmans

The pulse-length dependence of the laser-induced demagnetization has been


investigated systematically for pump pulses from 40 fs up to 7 ps [118]. It was
demonstrated that all M(t) curves could be reconstructed by simply convoluting the
shortest-pulse measurement with the respective pulse profiles. This behavior speaks
again in favor of a simple thermodynamic picture in which little role is played by
highly excited electrons and the laser field itself.
Roth et al. have presented a detailed study of laser-induced demagnetization of
a Ni thin film as a function of temperature (from 80 K up to 480 K), presented
in Fig. 16a, and laser fluences ranging up to almost complete demagnetization;
see Fig. 14b [85]. Upon increasing temperature, they observed a transition from a
fast demagnetization followed by a remagnetization to a two-step demagnetization
process, where after a fast decay, the magnetization continues to decrease at a slower
rate. The behavior at low and high temperatures has been dubbed type-I and type-
II, resp., as will be discussed in more detail below. Interestingly, the authors found
that all their M(t) traces could be described by a single set of parameters – most
specifically a single value of the spin-flip probability asf independent of temperature
and fluence. This way, they established a value asf in excellent agreement with the ab
initio estimates calculated by Carva et al., which range from 0.04 to 0.10, depending
on calculational details [119]. Despite this agreement, more work is still needed to
assess the relative contribution of e-p and e-e scattering to the demagnetization.
In the experiments by Roth et al., a characteristic slowing down is observed
in the τM versus ambient temperature and τM versus fluence dependence when
approaching full quenching. Such a slowing down at higher temperature and
fluence is a very characteristic behavior and was also reported for Ni and Co
in Ref. [103]. The same characteristic behavior was reported in a comparative
study on laser-induced demagnetization in Co films compared to Co/Pt multilayers
[120]. Measuring at a whole range of fluences, it was observed that τM for Co/Pt
(40–110 fs) is consequently below that for Co (50–300 fs); see Fig. 16b. Fitting the
trends with the M3TM revealed a fourfold increase of asf for the Co/Pt multilayers,
which was assigned to increased spin-orbit scattering with the heavy Pt atoms.
Following up the pioneering work by Vaterlaus [63], magnetization dynamics of
gadolinium has received intense interest, and important contributions have come in
particular by Bovensiepen and co-workers, using a variety of refined experimental
techniques, including MSHG [121], TR-MOKE [122], fs-XMCD [75], and
time-resolved angle-resolved photoemission spectroscopy [123]. Gadolinium is
an interesting material with a large local magnetic moment carried by the 4f
electrons and a much smaller moment due to an induced splitting of the itinerant 5d
bands, adding up to a total magnetic moment of 7.55 μB . Femtosecond excitation
could well cause strongly nonequilibrium dynamics, in which 4f and 5d magnetic
moments do not decay simultaneously. However, comparison of TR-MOKE and fs-
XMCD data, predominantly probing 5d and 4f , respectively, revealed that already
after a few hundred fs the two subsystems display almost identical dynamics
indicative for their efficient mutual exchange scattering [122].
One of the most striking findings is that the demagnetization in Gd proceeds
in two steps. Apart from a slow component at ∼50 ps, similar to the claim in the
10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films 519

(a) (b)

(c)
1.0
0.9
0.8
M/M0

0.7
0.6
0.5

-50 0 50 100 150 200


Delay (ps)

Fig. 16 (a) Ambient temperature dependence of the laser-induced sub-ps demagnetization of a


15 nm Ni film at constant laser fluence. A transition from a type-I to type-II behavior is observed
when increasing the ambient temperature from 300 to 480 K. Figure taken from Ref. [85]. (b)
Demagnetization time τM as a function of the demagnetization amplitude, measured in a Co/Pt
multilayer and a Co film. Figure taken from Ref. [120]. (c) Time-resolved XMCD signals measured
on a 10 nm Gd film using 50 fs laser pump pulses and ps (open symbols) or fs (solid symbols)
X-ray probe pulses. The solid red curve is a fit to the data using the M3TM. (Data points taken
from Ref. [75] with kind permission of APS)

original work [63], fs-XMCD studies revealed an initial much more rapid decay
within the first ps that accounts for only a part of the total demagnetization [75];
see Fig. 16c. As mentioned before, this behavior has been denoted as type-II, in
contrast to ordinary type-I behavior with a single continuous decay as observed for,
e.g., Ni and Co [103]. A similar two-step demagnetization was also observed in the
fs-XMCD for ferromagnetic Tb films, with an identical 1 ps initial decay, but the
slower decay being a factor of 5 faster than for Gd. The initial faster component
was assigned to the presence of hot electrons that speed up spin-lattice interaction
during the first ps, in line with the 1 ps that is needed in both materials to equilibrate
the electron and phonon system. The differences between Gd and Tb were assigned
to the different orbital momenta in their respective ground state, being L = 0 and
L = 3, respectively.
The peculiar type-II behavior observed for Gd has also been addressed in the
context of the M3TM in Ref. [103]. Surprisingly, in that work it was found that
type-II dynamics is a natural outcome whenever τM becomes larger than τE . In
such a case, a single scattering mechanism naturally provides a two-step decay with
two different time constants. The first rapid decay occurs as long as the electron
temperature is still (far) above the lattice temperature. Once the electron and phonon
520 M. L. M. Lalieu and B. Koopmans

temperatures have converged, the demagnetization process continues at a slower


rate, albeit driven by exactly the same microscopic process as during the first
phase. Independent support for the feasibility of type-II dynamics has come from
a T -dependent study on Ni [85], presented in Fig. 16a. There it was shown that
for Ni, the magnetization dynamics slows down so significantly at higher starting
temperatures that τM becomes larger than τE . A transition from type-I to type-II
behavior in the very same material is then expected, as was experimentally observed
indeed. In the nickel experiment, it is clear that no second mechanism is active. This
argument could be used as a smoking gun for the Gd and Tb case, although final
conclusions will require further, more detailed studies.
A study trying to intuitively understand the wide variety in demagnetization
time as observed for different materials was reported by Müller et al. [76]. They
compared demagnetization of materials with a relatively low spin polarization P
(Ni and the Heusler alloy Co2 MnSi) with that of materials with almost a 100%
spin polarization at EF (Fe3 O4 , LSMO, and CrO2 ). For the low-P materials, a very
rapid demagnetization τM ∼ 1 ps was measured, whereas the high-P materials all
showed a demagnetization at a 100 ps time scale, in some cases displaying a type-II
behavior. Also other materials (several oxides, selenides and sulfides) as measured
in Ref. [77] were included in the analysis. All these materials are assumed to be
close to half-metallic, i.e., having P = 1, and all of them have indeed showed long
demagnetization times ranging from a few ps to 1000 ps. Thus, a relation between P
and τM was concluded on. This conclusion can be considered to be consistent with
a spin-flip scattering model, since a significant reduction of asf could be expected if
one of the two spin bands has a gap in its band structure around EF .
Another prediction from LLB, ALLG, and M3TM is provided by the inverse
proportionality between τM and the Gilbert damping constant α, cf. Eq. (45). In Ref.
[55], where measurements of τM and α on the very same Ni thin film were reported,
a very close quantitative agreement with Eq. (45) was found. However, several
studies trying to verify the trend predicted by the equation for series of samples
did not find a consistent picture. As an example, Radu et al. measured τM and α for
permalloy as a function of impurity doping for several dopants, Ho, Dy, Tb, and Gd
[124]. A consistent reduction of τM with increasing doping concentration and α −1
was not observed. The authors concluded that the M3TM (as well as the LLB and
ALLG) seems to be oversimplified for treating the case of 4f impurities.
Although within the two decades after the discovery of sub-ps demagnetiza-
tion by Beaurepaire our understanding has increased enormously, full consensus
about the detailed mechanisms is still not being reached. Development of new
experimental approaches to investigate laser-induced magnetization dynamics will
be essential to make further progress. Among the outstanding questions, it is still
intensively debated whether the Stoner or Heisenberg picture is the most appropriate
for describing the ultrafast magnetic processes. As to this question, time- and spin-
resolved photoelectron spectroscopy provides additional insight by showing how
spin-polarized bands are evolving after laser excitation [125, 126]. Moreover, in
most descriptions of the magnetization dynamics, the spin system is approximated
to be internally in thermodynamic equilibrium. Just like with the electron and lattice
10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films 521

system, we know that this may be a poor approximation at a 100 fs time scale.
Indeed, experiments using time- and energy-resolved photoelectron spectroscopy
have resolved that at ultimate short time scales, one cannot treat the spin polarization
of the conduction band as a uniformly varying rigid system any longer [127].
However, whether such nonequilibrium effects significantly affect the main channel
of angular momentum transfer is yet unknown.

All-Optical Switching of Magnetization

The field of femtomagnetism gained a new boost by the discovery of all-optical


switching of magnetization in 2007 [5]. All-optical switching (AOS) describes
the reversal of the magnetization in a magnetic material by laser-pulse excitation,
without the use of any additional stimulus (e.g., an external magnetic field). The
discovery of AOS came entirely unexpected and gained much attention because the
optical switch is both fast (picosecond time scale) and energy-efficient and thus
shows high potential to be used in future magnetic memory applications. With
such applications in prospect, as well as a drive for fundamental knowledge, the
observation of AOS initiated a field of research that rapidly developed, pursuing full
comprehension of the mechanism underlying the AOS. In the following, a selected
overview of the research performed in this pursuit, and the resulting status quo is
given.

All-Optical Switching in Ferrimagnetic Alloys


The first demonstration of AOS was done using a magnetic thin film made of a
GdFeCo alloy [5]. This is a rare-earth transition-metal (RE-TM) alloy, which has
two distinct magnetic sublattices that are occupied by the Gd (RE) and FeCo (TM)
atoms. The alloy is a ferrimagnetic material, in which the two (ferro)magnetic sub-
lattices are coupled antiferromagnetically, and a (net) spontaneous magnetization
is maintained due to unequal sublattice magnetizations (typically at room tem-
perature). Figure 17a presents the pioneering measurement performed by Stanciu
et al., demonstrating AOS in a 20 nm thick GdFeCo film using single circularly
polarized femtosecond (fs) laser pulses [5]. The figure shows the out-of-plane
magnetization in the GdFeCo layer, initially containing both an up (white) and
down (black) domain. After sweeping a pulsing laser beam across the sample,
once with left (σ − ) and once with right (σ + ) circularly polarized laser pulses,
it can be seen that domains of reversed magnetization are “written” for specific
helicity and magnetization direction combinations. In other words, the final state
of the magnetization in these domains is determined by the helicity of the laser
pulse. Using these observations, the AOS was explained as a two-step process.
First the spin system is heated close to the Curie temperature, whereafter a light-
induced effective field, attributed to the IFE [61, 128, 129] (section “Conceptual
Introduction”), was assumed to switch the magnetization. The direction of the
effective field is set by the helicity of the light, making the AOS a helicity-dependent
process.
522 M. L. M. Lalieu and B. Koopmans

(a) (b)
100

Normalized XMCD (%)


Gd
50

–50
Fe

–100

–1 0 1 2 3
Pump–probe delay (ps)

Fig. 17 (a) Pioneering measurement demonstrating single-pulse all-optical switching in a


GdFeCo alloy, showing the out-of-plane magnetization in the layer, initially containing both an
up (white) and down (black) domain. Using both left (σ − ) and right (σ + ) circular polarized laser
pulses, it was demonstrated that the magnetization in the exposed area could be reversed and that
the final direction of the magnetization was determined by the helicity of the light. (Figure taken
from Ref. [5] with kind permission from APS). (b) Time-resolved element-specific x-ray magnetic
circular dichroism measurements on a GdFeCo thin film, showing the magnetization of the Gd
and Fe sublattices in time after fs laser-pulse excitation. (Figure taken from Ref. [6] with kind
permission from Springer Nature)

In the years that followed, the importance of both the laser-induced heating
and helicity effects for the AOS was under debate [130, 131, 132, 133, 134, 135].
Macrospin Landau-Lifshitz-Bloch [132] and atomistic Landau-Lifshitz-Gilbert
[134] calculations showed that AOS could indeed be achieved via a laser-induced
effective-field pulse (in combination with the demagnetization). However, apart
from the high amplitude of the magnetic field pulse needed for the switch (≈20 T),
the induced magnetic field needed to persist in the material for a much longer time
than the duration of the laser pulse itself. This resulted in the notion of some kind
of helicity memory mechanism in the RE-TM system [133, 135].
More clarity came when it was observed that the magnetization in GdFeCo can be
switched using single linearly polarized laser pulses [7], proving the (single-pulse)
switching mechanism to be a purely thermal process, which was simultaneously
established theoretically as well [136]. Moreover, it was demonstrated that the
helicity dependence presented in Fig. 17a was the result of magnetic circular dichro-
ism (MCD) [137], which describes the difference in absorption for left and right
circularly polarized light in magnetic materials (section “Basics of Magneto-Optics
from a Macroscopic Perspective”). In the case of AOS, there is a minimum amount
of energy that needs to be absorbed for the switch. Therefore, the MCD opens up a
fluence window where the AOS is helicity-dependent.
Insight into the thermal single-pulse switching mechanism was actually already
observed earlier using time-resolved XMCD measurements [6] (section “MO
Spectroscopy”). In these measurements, the magnetizations of the individual Gd
and Fe sublattices in a GdFeCo thin film were measured in time after fs laser-
pulse excitation and in the presence of an external magnetic field. The result of
10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films 523

the measurement is presented in Fig. 17b, where the normalized XMCD signal of
both magnetic sublattices is plotted as a function of pump-probe delay. The observed
behavior is typically divided into three subsequent phases: (i) the magnetization of
both sublattices rapidly demagnetizes after the laser-pulse excitation. The magneti-
zations on the two sublattices quench at different rates, with the Fe magnetization
demagnetizing much faster and reaching zero while the Gd is still demagnetizing.
(ii) As a result of exchange scattering, angular momentum is transferred from the
demagnetizing Gd to the Fe sublattice, resulting in a buildup of Fe magnetization
along the Gd magnetization direction. This results in a transient ferromagnetic-like
state in the ferrimagnet. (iii) Finally, the system cools down on a longer time scale.
With the relatively large magnetization buildup in the Fe, the antiferromagnetic
exchange coupling drives the switch of the Gd magnetization. Eventually, the system
recovers to the ferrimagnetic state, having its magnetization reversed with respect
to the initial state and demonstrating the AOS.
Soon it was concluded that – in agreement with the previous scenario – there are
three key ingredients necessary to facilitate the AOS: (i) two magnetic sublattices;
(ii) an antiferromagnetic exchange interaction between the two sublattices, resulting
in an antiparallel alignment of the sublattice magnetizations; and (iii) a difference
in demagnetization rate for the two sublattices. This was verified theoretically by
introducing these ingredients in various models, including a general theoretical
framework [136], atomistic spin models [6,7,138,139], and the M3TM [140], which
all demonstrated the capability of thermal single-pulse switching. Experimentally,
the range of materials for which AOS was found rapidly expanded, including
different RE-TM alloys, multilayers, heterostructures, and RE free synthetic-
ferrimagnetic heterostructures [141, 8, 142]. The common features in all these
material systems are the previously mentioned needed ingredients, supporting their
necessity. The importance of the helicity of the light, however, was not elucidated
for some of the systems.

All-Optical Switching in Ferromagnetic Systems


With the AOS mechanism and its requirements believed to be well understood, the
observation of AOS in ferromagnetic thin films, multilayers, and granular media
came as a surprise [9]. These ferromagnetic systems have a single magnetic layer,
thus lacking all of the ingredients previously considered necessary. Therefore, the
observation questioned the validity of the established understanding of the AOS
mechanism. Consequently, the inverse Faraday effect as a driving mechanism was
reconsidered. For instance, by adding an effective field pulse (as well as dipolar
interactions) to the M3TM discussed earlier (section “Theories for Femtosecond
Demagnetization”), the AOS images obtained in experiments on ferromagnets
could be reproduced [9, 143]. Additionally, ab initio calculations on the inverse
Faraday effect predicted the high field amplitudes required to model the AOS
[144]. Apart from the need of a deeper insight in the switching mechanism on a
fundamental level, the discovery gave the research field an additional boost since
the ferromagnetic materials showing AOS were the same systems already heavily
used in the field of spintronics. This observation could thus ease the integration of
AOS in future spintronic devices.
524 M. L. M. Lalieu and B. Koopmans

The inconsistency between the AOS mechanism resolved for the GdFeCo alloys
and the observation of AOS in ferromagnetic systems was reconciled by the
discovery that there are two different AOS mechanisms at play [145]. For the
GdFeCo system, it was confirmed that the AOS is indeed a single-pulse helicity-
independent switch, corresponding to the earlier discussed thermal single-pulse
switching mechanism. For both a ferromagnetic Pt/Co/Pt multilayer and a ferrimag-
netic TbCo alloy, however, no single-pulse AOS was found. For these systems, a
helicity-dependent multiple-pulse mechanism was revealed. A measurement of this
switching mechanism in the Pt/Co/Pt multilayer is shown in Fig. 18a. The figure

(a)

(b) -
M

1 Hz M 15 µm
No pulse 100 pulses 600 pulses 3,600 pulses
Fig. 18 (a) Measurement of the multiple-pulse AOS switching mechanism in a Pt/Co/Pt sample,
showing the normalized magnetization (measured as the anomalous Hall voltage VHall ) as a
function of time, while simultaneously the structure is exposed to femtosecond laser pulses which
are either left circular (σ − ), right circular (σ + ), or linearly (π ) polarized. (Figure taken from
Ref. [145] with kind permission from APS) (b) Measurement demonstrating the helicity-dependent
laser-pulse-induced domain wall motion. The direction of the domain wall motion is set by the
helicity of the laser pulses, causing either the up (light) or down (dark) magnetized domain to grow
in size. (Figure taken from Ref. [146])
10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films 525

shows the normalized magnetization (measured by the anomalous Hall voltage


VHall ) as a function of time, while simultaneously the structure is exposed to fs laser
pulses which are either left circularly (σ − ), right circularly (σ + ), or linearly (π )
polarized. As can be seen in the left panel, the magnetization in the Co layer is lost
during the first few laser pulses, independent of the polarization of the light. In the
case of the circularly polarized light, the magnetization recovers during subsequent
laser pulses (right panel), where the direction of the recovered magnetization is set
by the helicity of the light. No magnetization recovery is observed for the linearly
polarized laser pulses. This measurement clearly showed the helicity dependence in
the multiple-pulse AOS mechanism.
For the ferromagnetic thin films, the helicity dependence was proposed to
arise from laser-pulse-induced domain wall (DW) motion, which on itself was
experimentally demonstrated in Co/Pt multilayers [146, 147]. As presented in
Fig. 18b, the latter experiments showed that an up-down DW (i.e., a DW separating
an up (light) from a down (dark) magnetized domain) moves in opposite direction
when exposed by a train of left (σ − ) or right (σ + ) circularly polarized laser pulses.
In the case of the multidomain state generated by the first few laser pulses (left
panel Fig. 18a), the helicity-dependent DW motion causes the domains with a
certain magnetization direction to expand, causing the magnetization in the exposed
area to “grow” into the direction set by the helicity of the laser pulses. Both the
earlier discussed inverse Faraday effect and the magnetic circular dichroism were
suggested as a possible origin of the helicity-dependent DW motion [145,146,147].
Around the same time of the discovery of the two different switching mech-
anisms also came the observation of a domain-size criterion for the helicity-
dependent multiple-pulse AOS [148]. Using magnetic-layer-thickness-dependent
studies, it was revealed that the helicity-dependent AOS is observed when the
magnetic domain size during cooldown after the laser-pulse excitation is larger
than the laser spot size. This criterion can be seen as an extension to an earlier
found low-remanence criterion for ferrimagnetic alloys [149]. The latter showed
that the helicity-dependent multiple-pulse AOS in ferrimagnetic alloys is only
found when the remanent magnetization is below a specific value, resolving
the earlier observation that helicity-dependent AOS was only found when the
magnetization compensation temperature was close to room temperature [8]. Both
criteria demonstrate the need of a low dipolar energy to prevent (small) domain
formation during the multiple-pulse AOS.
The AOS in the (high-anisotropic) granular media was also found to be a
helicity-dependent multiple-pulse process [150]. However, a different (stochastic)
cumulative switching mechanism was proposed, based on a helicity-dependence of
the magnetization switching probability of the single-domain grains [151, 152, 150,
153]. This helicity-dependence was again linked to both the inverse Faraday effect
and the magnetic circular dichroism.

New Directions in All-Optical Switching


With the integration of AOS in future memory applications in mind, the research
on AOS shifted back toward the GdFeCo alloys because it was the only material
526 M. L. M. Lalieu and B. Koopmans

system for which the single-pulse AOS mechanism was found, which is needed
for fast future spintronics. Realizing that the AOS is a purely thermal process,
it was found that the switch could also be obtained using ultrafast hot-electron
pulses [154, 155] (analogous to the hot-electron pulse-induced demagnetization
discussed in section “Key Observations in Laser-Induced fs Demagnetization”).
These hot-electron pulses were generated by laser-pulse excitation of the Pt layer
in a Pt/NM/GdFeCo multilayer, whereafter the hot-electron pulse flows through the
thick nonmagnetic conductive spacer (NM) toward the GdFeCo layer. There, the
hot-electron pulse will deposit its energy and induce the magnetization reversal.
Later, using a photoconductive switch, it was demonstrated that the fast switch can
also be induced by a picosecond charge-current pulse [156].
Alongside the investigation on the different modes of stimuli for the thermal
AOS, research was done trying to achieve the single-pulse AOS in other material
systems that are more relevant for spintronic applications. For instance, it was
demonstrated that a ferromagnetic layer can be switched using single laser pulses
when it is deposited on top of a GdFeCo layer, in which case the ferromagnet
switches along with the GdFeCo alloy due to the exchange coupling between the
two layers [157]. A similar mutual AOS can be achieved when the GdFeCo layer
and the ferromagnet (FM) are separated by a nonmagnetic (NM) spacer layer to
create a GdFeCo/NM/FM spin-valve structure. In this case, the single-pulse AOS in
the GdFeCo alloy can be transferred to the ferromagnet by the spin current generated
by the Gd sublattice upon laser-pulse excitation [158, 159].
A different approach was taken, both theoretically [162] and experimentally
[160], by using a synthetic-ferrimagnetic multilayer which mimics the earlier
discussed properties of the GdFeCo alloy that are needed for the thermal single-
pulse AOS. Figure 19a shows the magnetization of a Pt/Co/Gd stack, toggling up
(light) and down (dark) upon excitation with subsequent fs laser pulses [160]. This
toggling behavior is characteristic for the thermal single-pulse AOS. The use of
such synthetic-ferrimagnetic multilayer allows for easy fabrication and (interface)
engineering. Moreover, the specific Pt/Co/Gd stack possesses a perpendicular
magnetic anisotropy (PMA) [160], large spin-orbit torques (spin Hall effect), and
a sizable interfacial Dzyaloshinskii-Moriya interaction (iDMI) resulting in chiral
Neél domain walls [161], making the structure an ideal candidate to facilitate the
integration of AOS with spintronic devices such as the racetrack memory [163]. The
potential of the stack for spintronic integration is demonstrated in Fig. 19c, showing
field-free “on-the-fly” AOS in a Pt/Co/Gd (micron-sized) racetrack [161]. The figure
shows the normalized anomalous Hall signal (∝ Mz ) measured in the Hall cross
located at one end of the racetrack, as illustrated in Fig. 19b. The measured signal
registers magnetic domains passing the cross shortly after they are written at the
other side of the racetrack using thermal single-pulse AOS (red dotted lines). The
domains are transported through the wire as soon as they are written by an electrical
current that is sent continuously through the wire, combining the spin Hall effect in
the heavy-metal Pt seed layer, the PMA in the magnetic layer, and the chiral Neél
walls for coherent and efficient domain wall motion.
Additionally, recent experiments have demonstrated that the DW velocity
in synthetic-antiferromagnetic [164] and (synthetic) ferrimagnetic [165, 166]
10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films 527

Fig. 19 (a) Kerr microscope 200 um


image of the magnetization in
a Pt/Co/Gd stack after
laser-pulse excitation. The 1 2 3 4 5
labels indicate the number of
subsequent linearly polarized
laser pulses the spot is
exposed to. Light and dark
6 7 8
regions represent the up and (a) 9 10
down (out-of-plane)
magnetization directions,
respectively. Figure taken LI 1 LI 2
from Ref. [160]. (b,c)
Measurement demonstrating
Ga+
“on-the-fly” AOS in a
J J
(micron-sized) Pt/Co/Gd
racetrack. (Figure taken from
Ref. [161] with kind (b)
permission from APS)
1.0
LI 2 (norm.)

0.5

0.0

-0.5

-1.0
0 5 10 15 20 25 30 35 40 45 50 55 60
(c) Time (s)

systems dramatically increases when having a compensated magnetization


or angular momentum, respectively. Combined with the easy fabrication and
engineering of the synthetic-ferrimagnetic multilayers used for the single-pulse
AOS, this shows great prospect for future integrated photonic memory applications.

Laser-Pulse-Excited Spin Currents

Around the same time of the first observation of all-optical switching, it was also
discovered that the fs laser-pulse excitation of ferromagnetic layers can induce a
nonlocal transfer of angular momentum [10]. In other words, it was demonstrated
that a current of spin polarized electrons, i.e., a spin current, can be generated upon
fs laser-pulse excitation of a ferromagnetic thin film. As was the case for all-optical
switching, part of the excitement over the observation derived from its potential use
in future spintronic applications. Within the field of spintronics, spin currents are
heavily used to manipulate magnetic information in (future) data storage devices,
e.g., to write data in magnetic random access memory [167] or transport data in
the earlier introduced magnetic racetrack memory [163]. Conventionally, these spin
currents are generated electronically on a nanosecond time scale. The manipulation
of the magnetization can be pushed to the ultrafast time scale by using the optically
528 M. L. M. Lalieu and B. Koopmans

generated spin currents, which operate at the sub-picosecond time scale. On a more
fundamental level, the discovery of the fs laser-pulse-excited spin current sparked
a debate about the underlying mechanism of ultrafast demagnetization, discussing
the importance of both local and nonlocal dissipation of angular momentum. In the
following, a selected overview of the research performed on the optically generated
spin currents and their ability to control the magnetization in magnetic thin films is
given.

Optically-Induced Spin Transfer


The presence of laser-pulse-excited spin currents was first demonstrated in a
collinear magnetic bilayer, consisting of two identical out-of-plane magnetized FM
layers that were separated by a nonmagnetic conductive spacer [10]. As illustrated
in Fig. 20a, the magnetizations in the FM layers can either be in a parallel (P)
or an antiparallel (AP) alignment. Upon laser-pulse excitation, both FM layers
demagnetize, and angular momentum is exchanged between the two layers by
the spin currents that are generated in each layer and are injected into the other
layer. The spin current leaving each layer is polarized along the magnetization
in the concerning layer, as shown in Fig. 20a. As a result, in the case of a
parallel (antiparallel) alignment, majority (minority) spins are injected in each layer
after laser-pulse excitation, which increases (decreases) the magnetic moment in
the layer and thereby hinders (assists) the demagnetization. The measured effect
of the transfer of angular momentum on the demagnetization in either layer is

(a) (b)
t Ru = 0.4 nm
Normalized Δθ/θ

(c) H probe
pump Fe Au AP
P
MgO(001)

0 0.5 1.0
M mSHG Delay (ps)
Fig. 20 (a) Illustration of the magnetic collinear bilayer used to demonstrate the presence of laser-
pulse-excited spin currents. (b) Demagnetization traces measured with the magnetic bilayer in the
parallel (P) and antiparallel (AP) configuration, showing an increase in demagnetization speed and
amplitude for the AP alignment with respect to the P alignment. ((a,b) Figure taken from Ref.
[10] with kind permission of Springer Nature) (c) Illustration of the second-harmonic generation
pump-probe measurement on a Fe/Au bilayer, directly measuring the laser-induced spin current
injected from the Fe layer into the nonmagnetic Au layer. (Figure taken from Ref. [168] with kind
permission from APS)
10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films 529

shown in Fig. 20b, which indeed shows that both the speed and amplitude of the
demagnetization is increased for the antiparallel alignment with respect to the
parallel alignment. Performing similar measurements on a bilayer with an insulating
spacer showed no difference in the demagnetization for the parallel and antiparallel
configurations, confirming the presence of angular momentum exchange by the
laser-pulse-excited spin currents in the case of the conducting spacer.
Soon after the discovery, a first model on the generation mechanism of the
optically excited spin current was developed [11, 169]. In this model, a spin current
is generated as a result of spin-dependent lifetimes and velocities of the excited
hot electrons. As a specific example, the hot majority electrons in Ni have a longer
inelastic lifetime as compared to minority electrons [τ0,↑ > τ0,↓ , Eq. (40)], and
thereby will dominate the spin current. The spin current is considered to propagate in
the “superdiffusive” regime, starting in the ballistic regime just after excitation and
relaxing to normal diffusion on a longer time scale. Moreover, it was claimed that
the demagnetization measured in thin nickel films could be explained purely based
on this nonlocal transfer of angular momentum away from the probed area, starting
the discussion on the local versus nonlocal dissipation of angular momentum as the
driving force of fs laser-induced demagnetization.
Even though the occurrence of optically induced spin currents was demonstrated
in experiments such as shown in Fig. 20a, b, determining their relative contribution
to sub-ps demagnetization as compared to local dissipation of angular momentum
turned out to be a challenge. While some experiments were explained using the
optically excited spin current as the dominant source of demagnetization [170, 171,
172, 173], other demagnetization studies showed no sign of its contribution [93].
An increasing amount of experiments, however, demonstrated that both the local
and nonlocal angular momentum dissipation mechanisms are of importance and
that the relative contribution of either is dependent on the precise material system
[34,29,174,46]. Using element-specific XMCD measurements, even a laser-induced
enhancement of the Fe magnetization in a Fe/Ru/Ni multilayer was claimed [170,
33, 34]. The enhancement to values up to 10% above the saturation magnetization
was attributed to the spin current generated in the (parallel) Ni layer being absorbed
in the Fe layer. This observation, however, could not be reproduced using different
measurement techniques, i.e., by XMCD measurements at a different absorption
edge of the Fe and Ni [174], or using the earlier discussed layer-specific MOKE
technique [28] (see section “Layer-Specific MOKE”).
The first direct detection of the optically excited spin current was done using a
Fe/Au bilayer [168]. An illustration of the measurement is shown in Fig. 20c. A spin
current was generated in the Fe layer by laser-pulse excitation at the Fe side, which
was injected into the (thick) nonmagnetic Au layer and propagated toward the outer
Au surface. Using a probe pulse incident from the Au side, the spin current was
measured at the Au surface via magnetization-induced second-harmonic generation
(MSHG) (see section “MO Spectroscopy”). Later, the spin accumulation resulting
from this laser-pulse-excited spin-current injection into the nonmagnetic layer in a
FM/NM bilayer was also measured using time-resolved MOKE [13, 175] and using
the layer-specific MOKE technique [46].
530 M. L. M. Lalieu and B. Koopmans

With the increasing amount of research on the optically excited spin currents,
more mechanisms for spin current generation were proposed. At least four different
mechanisms have been identified. (i) As conjectured in the original work by
Malinowski [10], an optically generated spin-polarized distribution of ballistic (hot)
electrons can act as a source of spin current. (ii) Adding inelastic scattering events,
one arrives at the earlier discussed superdiffusive spin current model [11, 169]. (iii)
Beyond the superdiffusive mechanism, assuming all electrons to be thermalized,
a laser-induced temperature gradient across a ferromagnetic layer may still result
in a longer-lived spin current generation at the FM/NM interface due to the spin-
dependent Seebeck effect [176]. This concept was extended to the nonthermal
regime in Ref. [177]. (iv) Also the rapid change of M itself has been proposed
as a source. More specifically, the angular momentum conservation of the electron-
magnon coupling causes the demagnetization to generate spin-polarized electrons
at a rate of −dM/dt, which propagate into the neighboring layers. This process was
modeled in the diffusive regime by Choi et al. [13]. Strongly related, but within
the perspective of an s-d model, sudden laser-driven demagnetization will induce a
temporal splitting of the s-electron chemical potential, which also act as a source of
spin current into neighboring layers [113].

Optical Spin-Transfer Torque


Alongside the investigation on the underlying mechanism, research focused on the
optically generated spin current itself and its ability to manipulate the magnetization
in a second FM layer on an ultrafast time scale [12, 13, 176, 178, 179]. As an
example, experiments were performed using noncollinear bilayers, as illustrated in
Fig. 21a. In these experiments, the optically excited spin current generated in an out-
of-plane magnetized layer (FM1), called the emitter, travels through a conductive
spacer layer and is injected into an in-plane (transversally) magnetized layer (FM2),
called the collector. The injected transverse spins are absorbed in the collector,
which results in a spin-transfer torque (STT) on its magnetization [180], canting
the magnetization toward the direction of the injected spins.
One way to measure the magnetization canting is by performing such an
experiment in the presence of an external magnetic field that is aligned parallel to
the (initial) direction of the magnetization in the collector. After the initial canting
of the magnetization due to the absorbed spin-current pulse, the magnetization is no
longer aligned with the applied magnetic field, resulting in a damped precessional
motion of the magnetization around the applied field direction [56] (see Eq. (35)).
An example of such a precession measurement for a [Co/Ni]4 /Co/Cu/Co system is
presented in Fig. 21b. The figure shows the measured Kerr rotation as a function of
time after the fs laser-pulse excitation (demagnetization signal is subtracted), mea-
sured with a polar TR-MOKE setup (section “Measuring Ultrafast Magnetization
Dynamics”), and for two different configurations of the noncollinear magnetiza-
tions. The measured dynamics is a superposition of two precessions, which are
individually shown by the red and blue solid curves for the top measurement. The
blue curve corresponds to the precession of the out-of-plane magnetization in the
[Co/Ni]4 emitter (blue arrow) and originates from a laser-induced change in the
10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films 531

(a) Pt FM1 Cu FM2

Heat current
M spin current M

(b) 0.2
0.1
(mdeg)

0.0

–0.1
0.1
Δ Kerr

0.0

–0.1
0 250 500 750 1,000
Delay (ps)

(c) 8

6
Efficiency (%)

0
0 1 2 3 4 5 6
Co thickness (nm)

Fig. 21 (a) Illustration of a noncollinear bilayers used to measure the laser-induced spin-transfer
torque. (Figure taken from Ref. [13] with kind permission from Springer Nature) (b) Time-resolved
measurements of the out-of-plane component of the magnetization for two different configurations
of the noncollinear bilayer. (Figure adapted from Ref. [12] with kind permission from Springer
Nature) (c) Efficiency as a function of the collector layer thickness, with the efficiency defined as
the ratio of spins absorbed by the collector to the spins lost during demagnetization in the emitter.
(Figure taken from Ref. [179] with kind permission from APS)
532 M. L. M. Lalieu and B. Koopmans

anisotropy [12, 56] (as discussed in section “Conceptual Introduction”). The red
curve, however, corresponds to the precession of the in-plane magnetization in the
Co collector (red arrow) and is driven by the optical STT mechanism demonstrated
in Fig. 21a. Comparing the dynamics for the two different configurations shows that
the phase of the magnetization precession in the collector is set by the orientation of
the magnetization in the emitter (i.e., the polarization of the generated spin current),
as is expected for the laser-induced STT mechanism.
This laser-induced STT in the noncollinear bilayer can be used as a tool to probe
the optically excited spin current and has been used to investigate the transport of
the laser-pulse-excited spin current through different spacer layers [12], the optical
excitation of spin currents in ferrimagnetic alloys [159], and the dependence of the
optical spin current generation on the emitter thickness [179]. The latter showed that
the full thickness of the emitter (an out-of-plane magnetized [Co/Ni]N multilayer)
contributes to the generated spin current (at least up to 3.4 nm), which is consistent
with recently performed theoretical calculations using an extended superdiffusive
spin current model [181] but contradicts an earlier observed optical spin-emission
region in Fe of ≈1 nm [182]. Additionally, the absorption of the optically excited
spin current in the collector was investigated using a wedge-shaped collector layer.
The result of this measurement is presented in Fig. 21c. The figure shows the
efficiency as a function of the Co collector thickness, with the efficiency defined as
the ratio of spins absorbed by the collector to the spins lost during demagnetization
in the [Co/Ni]4 emitter. It can be seen that the spin current is absorbed very locally
near the injection interface, i.e., 90 % of the transverse spins are absorbed within
the first ≈2 nm of the collector layer.

Optical THz Spin Wave Excitation


As it turns out, the very local absorption of the short and intense optically generated
spin current pulse near the injection interface provides ideal conditions for the
excitation of higher-order (sub)THz standing spin waves in the collector [178, 179].
This is demonstrated for a noncollinear bilayer with a 5.5 nm thick Co collector
layer in Fig. 22a. The measured polar TR-MOKE signal shows both the first-order
standing spin wave (0.55 THz) as well as the uniform precession (≈10 GHz). As
illustrated in the inset of the figure, the higher-order THz spin waves are excited via
the creation of a strong gradient in the canting angle of the magnetization, created by
the very local absorption of the optically generated spin current near the interface.
This highly nonequilibrium magnetization state leads to (damped) standing spin
waves, which are driven by the exchange interaction and can be excited without the
presence of an externally applied field. The frequency of the standing spin waves is
determined by their wavelength, which in turn is set by the thickness of the collector
and reaches above 1 THz for a 3 nm thick Co layer [179].
In the measurement presented in Fig. 22a, only the first-order standing spin
wave is observed. Higher-order modes were observed in a different measurement
performed on a noncollinear bilayer with two (perpendicularly aligned) in-plane
magnetized layers and with a thicker collector (a 14 nm thick Fe layer) [178].
The result of that measurement is presented in Fig. 22b. In this figure, the left
10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films 533

(a) 4.0

3.5
MOKE signal (arb. unit.)

3.0

2.5

2.0 0.55 THz


1.5

1.0

0.5

0.0
10.23 GHz
-0.5
0 5 20 40 60 80 100 120 140 160 180 200
Delay (ps)

(b) 1.2
(c)
×20 n =4

n =3
Frequency (THz)

0.8 n =2

n =1

n =0
0.4

DFe = 280 meV·Å2


0
0 1 2 3 4 5 6 d = 14 nm
Eigenmode number n = k d /π

Fig. 22 (a) Measurement of the (sub)THz standing-spin-wave excitation using a noncollinear


bilayer with a 5.5 nm thick Co collector layer. The figure shows the polar TR-MOKE signal
measured as a function of time after the laser-pulse excitation. Both the first-order standing spin
wave (0.55 THz) as well as the uniform precession (≈10 GHz) are visible. The inset shows the
gradient in the magnetization within the collector layer after the optical excitation and the resulting
uniform and first-order standing spin wave. Figure taken from [179]. (b) Left panel: Fourier
spectrum of the magnetization dynamics measured in a 14 nm thick Fe collector layer after laser-
pulse excitation. Right panel: calculated spin-wave dispersion, in which the dots represent the
frequencies corresponding to the different standing-spin-wave orders. There is a match between
the peaks in the Fourier spectrum and the calculated standing-spin-wave frequencies up to n = 4
(solid dots). (c) Illustration of the different standing-spin-wave orders. ( (b,c) Figure taken from
Ref. [178])

panel shows the Fourier spectrum of the magnetization dynamics measured in the
collector after laser-pulse excitation. The right panel displays the calculated spin
wave dispersion, in which the dots represent the frequencies corresponding to the
534 M. L. M. Lalieu and B. Koopmans

different standing-spin-wave orders. As can be seen, there is a match between the


peaks in the Fourier spectrum and the calculated standing-spin-wave frequencies
up to n = 4 (solid dots), demonstrating that standing spin waves up to the fourth-
order mode were excited in the noncollinear bilayer by fs laser-pulse excitation;
see also Fig. 22c. Recently, the higher-order standing spin wave excitation via
the ultrafast spin-transfer torque from an injected fs transverse spin current pulse
was reproduced using micromagnetic modeling [183]. Moreover, an experimental
design that allows for the lateral propagation of these laser-pulse-excited spin waves
was suggested. These findings show that in addition to its general importance in the
field of spintronics, the optically excited spin currents could also be of high potential
for future THz magnonics.

Conclusions and Outlook

At the end of this chapter, we briefly highlight some of the exciting trends in the
field and sketch promising routes for future research.
Magneto-optics is known for almost two centuries, and experimental approaches
including Kerr microscopy have well matured. Nevertheless, applications to new
systems and phenomena keep introducing new demands and opportunities. The
rise of thin film magnetism and the birth of spintronics in the last two decades
of the previous century have pushed applications of the magneto-optic Kerr effect
enormously. A similar boost has come from recent trends in spintronics, such as
novel device architectures based on magnetic domain walls that are driven by spin-
orbit torques, which has caused a revival of Kerr microscopy.
Another trigger pushing methodological developments has come from the new
field of opto-magnetism, requiring magneto-optical techniques with femtosecond
resolution. In many cases this can be easily established by pump-probe configura-
tions and measuring in a stroboscopic manner. But such an approach is of limited
value if stochasticity plays a role. This faces new challenges, in particular when
aiming at microscopy. Developments of single-shot wide-field magneto-optical
microscopy and other experimental approaches with combined spatial resolution
down to the nanometer scale and sub-ps time resolution are emerging and will most
probably grow in importance. An interesting case is provided by establishing depth
resolution, which could certainly be pushed beyond todays achievements.
As briefly discussed in this chapter, together with experimental progress in
using visible and near-infrared (laser) light for magneto-optical and opto-magnetic
experiments, rapid developments using other parts of the electromagnetic spectrum
have been witnessed. In this sense X-ray techniques are particularly noteworthy
because of their element specificity, as well as distinguishing spin and orbital
magnetism. Table-top implementations using high-harmonic generation will enable
much more widespread use and possibly new spectroscopic approaches. Finally, a
rapid growth of THz studies is being seen. They provide access to ac conductivity
of electrons and a complementary view on sub-ps magnetization dynamics and spin
currents.
10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films 535

In the field of opto-magnetism, a trend toward further refinement of the optical


manipulation will certainly continue. Prominent issues on the agenda for the
forthcoming decade will be improving the energy efficiency as well as spatial
resolution of all-optical switching. As to the latter, first attempts of using plasmonics
for further downscaling beyond the diffraction limit have been reported, but it will
take intensive research efforts to push this direction toward more robust control.
Another issue relates to exploring the ultimate data rates possible. Note that the fact
that we can switch magnetization by sub-ps laser pulses is scientifically of interest,
but for applications the data rate – determined by the minimum time duration
between two succeeding pulses to establish two independent switching processes –
will be the decisive factor.
Pushing all-optical manipulation (switching) of magnetic matter to a next level of
control will certainly rely on a better understanding of the underlying processes. The
basics of the single-pulse exchange-driven toggle mechanism are well understood
and can be qualitatively reproduced by different models. However, a number of
outstanding questions are indicating that our knowledge is far from complete.
It seems that Gd is an essential element to establish single-pulse AOS, but it
is not known yet why. Also, it has been claimed that spin transport may be of
crucial relevance for successful AOS, but its importance has yet to be confirmed.
More general, insight is lacking as to the relative role of electron-mediated spin
currents, exchange scattering, and magnon spin currents in the AOS process.
Beyond understanding the toggle-switching scenario, many open questions exist
with respect to the origin of helicity-dependent switching, whether it can be pushed
toward the single-pulse regime and about the role of the laser pulse length.
Beyond using rare-earth-transition ferromagnetic metal alloys and layered syn-
thetic ferrimagnets, it is of significant interest to explore other materials that
could display the same or similar phenomena, maybe – at least with respect to
some aspects – in a superior way. In this context, entirely different means of
nonthermal switching in magnetic dielectric materials has been successfully started
to be explored. These dielectrics provide a complementary route toward highly
efficient, low power switching. For sure the switching mechanism is by far superior
when it comes to dissipated laser power, but due to the low absorption, it needs
extremely high laser power to excite. Moreover, the specific materials may be less
easily integrated in devices, and future research should clarify its true potential.
Another highly interesting route is linked to the present explosion in the use of
antiferromagnets for spintronics. Ideas are emerging as to how antiferromagnets
can be switched using ultrashort laser pulses.
A particularly noteworthy trend that we signaled in this chapter is the merge of
magneto-optics, opto-magnetism, and spintronics. We foresee that this development
will further evolve. On the one hand, it provides and interesting route toward novel
applications in a More than Moore context (NB: More than Moore refers to a trend
in modern chip technology to embed new and complementary functionalities on-
chip, rather than further pushing miniaturization according to Moore’s law). On
the other hand, it provides a beautiful playground for fundamental research. The
demonstrated ability to generate spin currents by fs laser pulses at the nanometer
536 M. L. M. Lalieu and B. Koopmans

scale allows for a complementary view on spintronic concepts and in particular to


explore their ultimate limits. We emphasize that the recent finding that AOS can
be established in exactly the type of easy-engineerable ultrathin-layered systems
used for spintronics will further fuel the successful merge of opto-magnetism and
spintronics. In this context it is also worthwhile mentioning recent THz studies
on ultrafast magnetization dynamics, as another means of driving ac free-electron
(spin) currents.
Toward further understanding of the ultrafast laser-induced magnetization
dynamics, including AOS, an important role will inevitably be played by refinement
of theory and increase of numerical abilities. As to magneto-optics, ab initio density
functional schemes for calculating magneto-optical response have been successfully
introduced already by the end of the previous century, but the problem describing
femto-second opto-magnetism is orders of magnitude more complex. In the past
decade, a number of semiempirical theoretical frameworks have been developed
that can successfully reproduce – and sometimes even predict – a wide range of
experiments. However, it is only since very recently that fully ab initio approaches,
such as time-dependent density functional theory, have been able to make contact
with laser-excited femtomagnetism. While before 2010 it was still unimaginable to
include all involved degrees of freedom (electronic, lattice, and spin) with enough
accuracy in a time-dependent ab initio code, due to more efficient codes and increase
of computational power, promising breakthroughs are now being reported. It is
expected that such ab initio approaches will become of increasing importance to
guide the field further in the forthcoming decade.
Some final words are reserved for sketching the application potential of the
exciting phenomena discussed in the present chapter. Commercial use of magneto-
optical recording dates back to the mini-disk system in the 1990s. Despite high
expectations in the decade that followed, next generations of optical storage tech-
nology moved away from magneto-optics. However, a very exciting combination
of magnetic hard disk recording with optics is presently becoming more mature. In
heat-assisted magnetic recording (HAMR), an optical waveguide is integrated with
the magnetic write head. Laser light is focused to a 50 nm spot using plasmonic
tricks, to locally heat up the storage medium and lower its coercivity. This process
enables using harder magnetic media and thus pushes data storage densities. Very
excitingly, after 20 years of industrial development, this approach is now at the brink
of being launched commercially.
This success of HAMR might fuel expectations for other application using com-
binations of light and magnetism. Actually, ever since the first reports on ultrafast
laser-induced magnetization dynamics, and in particular AOS, it has been (maybe
overoptimistically) positioned as baring enormous potential for applications. Its
potential for recording yet has to be proven, but a particular appealing direction
is to consider applications beyond (disk-based) storage technology. Both magneto-
optics and opto-magnetism could be envisioned to be implemented in photonic
integrated circuits, to enrich their functionality. Magneto-optics has been considered
for realizing a compact on-chip optical isolator already for decades, albeit with little
success so far. However, first research initiatives have been established that aim at
10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films 537

a broader implementation of spintronic and magnetic functionalities – in particular


include AOS – in integrated photonics. It is a fascinating thought that one day the
merge of magneto-optics, opto-magnetism, and spintronics might lead to spin off in
such an applied photonic context.

Notes
1 An exception is provided by topological MO effects for crystals with a specific magnetic point
group symmetry with a chiral spin ordering [19].
2 Note that in the presence of spin-orbit coupling S and L are no good quantum numbers either, but

we might have expected the total J to be still conserved, but this is apparently not the case.

References
1. Faraday, M.: On the magnetization of light and the illumination of magnetic lines of force.
Trans. Roy. Soc. 5, 592 (1864)
2. Kerr, J.: On the rotation of the plane of polarization by reflection from the pole of a magnet.
Phil. Mag. 3, 321–343 (1877)
3. Mansuripur, M.: The Physical Principles of Magneto-Optical Recording. Cambridge Univer-
sity Press (1995)
4. Beaurepaire, E., Merle, J.-C., Daunois, A., Bigot, J.-Y.: Ultrafast spin dynamics in ferromag-
netic nickel. Phys. Rev. Lett. 76, 4250–4253 (1996)
5. Stanciu, C.D., Hansteen, F., Kimel, A.V., Kirilyuk, A., Tsukamoto, A., Itoh, A., Rasing, T.:
All-optical magnetic recording with circularly polarized light. Phys. Rev. Lett. 99, 047601
(2007)
6. Radu, I., Vahaplar, K., Stamm, C., Kachel, T., Pontius, N., Durr, H.A., Ostler, T.A., Barker,
J., Evans, R.F.L., Chantrell, R.W., Tsukamoto, A., Itoh, A., Kirilyuk, A., Rasing, T., Kimel,
A.V.: Transient ferromagnetic-like state mediating ultrafast reversal of antiferromagnetically
coupled spins. Nature 472, 205–208 (2011)
7. Ostler, T.A., Barker, J., Evans, R.F.L., Chantrell, R.W., Atxitia, U., Chubykalo-Fesenko, O.,
El Moussaoui, S., Le Guyader, L., Mengotti, E., Heyderman, L.J., Nolting, F., Tsukamoto, A.,
Itoh, A., Afanasiev, D., Ivanov, B.A., Kalashnikova, A.M., Vahaplar, K., Mentink, J., Kirilyuk,
A., Rasing, T., Kimel, A.V.: Ultrafast heating as a sufficient stimulus for magnetization
reversal in a ferrimagnet. Nat. Commun. 3, 666 (2012)
8. Mangin, S., Gottwald, M., Lambert, C-H., Steil, D., Uhlír, V., Pang, L., Hehn, M., Alebrand,
S., Cinchetti, M., Malinowski, G., Fainman, Y., Aeschlimann, M., Fullerton, E.E.: Engineered
materials for all-optical helicity-dependent magnetic switching. Nat. Mater. 13, 286–292
(2014)
9. Lambert, C-H., Mangin, S., Varaprasad, B.S.D.C.S., Takahashi, Y.K., Hehn, M., Cinchetti,
M., Malinowski, G., Hono, K., Fainman, Y., Aeschlimann, M., Fullerton, E.E.: All-optical
control of ferromagnetic thin films and nanostructures. Science 345, 1337–1340 (2014)
10. Malinowski, G., Della Longa, F., Rietjens, J.H.H., Paluskar, P.V., Huijink, R., Swagten,
H.J.M., Koopmans, B.: Control of speed and efficiency of ultrafast demagnetization by direct
transfer of spin angular momentum. Nat. Phys. 4, 855–858 (2008)
11. Battiato, M., Carva, K., Oppeneer, P.M.: Superdiffusive spin transport as a mechanism of
ultrafast demagnetization. Phys. Rev. Lett. 105, 027203 (2010)
12. Schellekens, A.J., Kuiper, K.C., de Wit, R.R.J.C., Koopmans, B.: Ultrafast spin-transfer
torque driven by femtosecond pulsed-laser excitation. Nat. Commun. 5, 4333 (2014)
13. Choi, G-M., Min, B-C., Lee, K-J., Cahill, D.G.: Spin current generated by thermally driven
ultrafast demagnetization. Nat. Commun. 5, 4334 (2014)
538 M. L. M. Lalieu and B. Koopmans

14. Zak, J., Moog, E.R., Liu, C., Bader, S.D.: Universal approach to magneto-optics. J. Magn.
Magn. Mat. 89, 107–123 (1990)
15. Jones, R.C.: A new calculus for the treatment of optical systems. J. Opt. Soc. Am. 31, 488–493
(1941)
16. Rzazewski, K., Boyd, R.W.: Equivalence of interaction Hamiltonians in the electric dipole
approximation. J. Mod. Opt. 51, 1137–1147 (2004)
17. Oppeneer, P.M.: Magneto-optical kerr spectra. In: Buschow, K.H.J. (ed.) Handbook of
Magnetic Materials: Volume 13, pp. 229–422. Elsevier Science (2001)
18. Bennett, H.S., Stern, E.A.: Faraday effects in solids. Phys. Rev. 137, A448–A470 (1965)
19. Feng, W., Zhou, X., Hanke, J-P., Guo, G-Y., Blügel, S., Mokrousov, Y., Yao, Y.: Topo-
logical magneto-optical effect and its quantization in noncoplanar antiferromagnets (2018)
arXiv:1811.05803.
20. Oppeneer, P.M.: Magneto-optical spectroscopy in the valence-band energy regime: relation-
ship to the magnetocrystalline anisotropy. J. Magn. Magn. Mat. 188, 275–285 (1998)
21. Krinchik, G.S., Artem’ev, V.A.: Magneto-optical properties of Ni, Co, and Fe in the ultraviolet
visible, and infrared parts of the spectrum. Zh. Eksp. Teor. Fiz. 53, 1901–1912 (1968)
22. Schellekens, A.J.: Manipulating spins. Ph.d. thesis, Eindhoven University of Technology
(2014)
23. Ebert, H.: Magneto-optical effects in transition metal systems. Rep. Prog. Phys. 59, 1665–
1735 (1996)
24. Hashimoto, S., Ochiai, Y., Aso, K.: Film thickness dependence of magneto-optical and
magnetic properties in Co/Pt and Co/Pd multilayers. J. Appl. Phys. 67, 4429–4431 (1990)
25. Koopmans, B., van Kampen, M., Kohlhepp, J.T., de Jonge, W.J.M.: Femtosecond spin
dynamics of epitaxial Cu(111)/Ni/Cu wedges. J. Appl. Phys. 87, 5070 (2000)
26. Koopmans, B.: Laser-induced magnetization dynamics. In: Hillebrands, B., Ounadjela, K.
(eds.) Spindynamics in Confined Magnetic Structures II, vol. 87, pp. 253–320. Springer
(2003)
27. Hamrle, J., Ferré, J., Nývlt, M., Višňovský, Š.: In-depth resolution of the magneto-optical kerr
effect in ferromagnetic multilayers. Phys. Rev. B 66, 224423 (2002)
28. Schellekens, A.J., de Vries, N., Lucassen, J., Koopmans, B.: Exploring laser-induced
interlayer spin transfer by an all-optical method. Phys. Rev. B 90, 104429 (2014)
29. Wieczorek, J., Eschenlohr, A., Weidtmann, B., Rösner, M., Bergeard, N., Tarasevitch, A.,
Wehling, T.O., Bovensiepen, U.: Separation of ultrafast spin currents and spin-flip scattering
in Co/Cu(001) driven by femtosecond laser excitation employing the complex magneto-
optical Kerr effect. Phys. Rev. B 92, 174410 (2015)
30. Khorsand, A.R., Savoini, M., Kirilyuk, A., Kimel, A.V., Tsukamoto, A., Itoh, A., Rasing, T.:
Element-specific probing of ultrafast spin dynamics in multisublattice magnets with visible
light. Phys. Rev. Lett. 110, 107205 (2013)
31. Thole, B.T., Carra, P., Sette, F., van der Laan, G.: X-ray circular dichroism as a probe of
orbital magnetization. Phys. Rev. Lett. 68, 1943–1946 (1992)
32. Carra, P., Thole, B.T., Altarelli, M., Wang, X.: X-ray circular dichroism and local magnetic
fields. Phys. Rev. Lett. 70, 694–697 (1993)
33. Mathias, S., La o vorakiat, C., Shaw, J.M., Turgut, E., Grychtol, P., Adam, R., Rudolf, D.,
Nembach, H.T., Silva, T.J., Aeschlimann, M., Schneider, C.M., Kapteyn, H.C., Murnane,
M.M.: Ultrafast element-specific magnetization dynamics of complex magnetic materials on
a table-top. J. Electron. Spectrosc. Relat. Phenom. 189, 164–170 (2013)
34. Turgut, E., La o vorakiat, C., Shaw, J.M., Grychtol, P., Nembach, H.T., Rudolf, D., Adam, R.,
Aeschlimann, M., Schneider, C.M., Silva, T.J., Murnane, M.M., Kapteyn, H.C., Mathias, S.:
Controlling the competition between optically induced ultrafast spin-flip scattering and spin
transport in magnetic multilayers. Phys. Rev. Lett. 110, 197201 (2013)
35. Pan, R-P., Wei, H.D., Shen, Y.R.: Optical second-harmonic generation from magnetized
surfaces. Phys. Rev. B 39, 1229–1234 (1989)
36. Koopmans, B., Janner, A-M., Jonkman, H.T., Sawatzky, G.A., van der Woude, F.: Strong bulk
magnetic dipole induced second-harmonic generation from C60 . Phys. Rev. Lett. 71, 3569–
3572 (1993)
10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films 539

37. Qiu, Z.Q., Bader, S.D.: Surface magneto-optic kerr effect. Rev. Sci. Instrum. 71, 1243–1255
(2000)
38. Allwood, D.A., Xiong, G., Cooke, M.D., Cowburn, R.P.: Magneto-optical Kerr effect analysis
of magnetic nanostructures. J. Phys. D Appl. Phys. 36, 2175–2182 (2003)
39. Schäfer, R.: Investigation of domains and dynamics of domain walls by the magneto-optical
Kerr effect. In: Kronmüller, H., Parkin, S. (eds.) Handbook of Magnetism and Advanced
Magnetic Materials. Wiley (2007)
40. Soldatov, I.V., Schäfer, R.: Selective sensitivity in kerr microscopy. Rev. Sci. Instrum. 88,
073701 (2017)
41. Buess, M., Höllinger, R., Haug, T., Perzlmaier, K., Krey, U., Pescia, D., Scheinfein, M.R.,
Weiss, D., Back, C.H.: Fourier transform imaging of spin vortex eigenmodes. Phys. Rev.
Lett. 93, 077207 (2004)
42. Neudert, A., McCord, J., Chumakov, D., Schäfer, R., Schultz, L.: Small-amplitude mag-
netization dynamics in permalloy elements investigated by time-resolved wide-field Kerr
microscopy. Phys. Rev. B 71, 134405 (2005)
43. Bigot, J.-Y., Guidoni, L., Beaurepaire, E., Saeta, P.N.: Femtosecond spectrotemporal
magneto-optics. Phys. Rev. Lett. 93, 077401 (2004)
44. Ju, G., Vertikov, A., Nurmikko, A.V., Canady, C., Xiao, G., Farrow, R.F.C., Cebollada, A.:
Ultrafast nonequilibrium spin dynamics in a ferromagnetic thin film. Phys. Rev. B 57, R700–
R703 (1998)
45. Dalla Longa, F.: Laser-induced magnetization dynamics. Ph.d thesis, Eindhoven University
of Technology (2008)
46. Hofherr, M., Maldonado, P., Schmitt, O., Berritta, M., Bierbrauer, U., Sadashivaiah, S.,
Schellekens, A.J., Koopmans, B., Steil, D., Cinchetti, M., Stadtmüller, B., Oppeneer, P.M.,
Mathias, S., Aeschlimann, M.: Speed and efficiency of femtosecond spin current injection
into a nonmagnetic material. Phys. Rev. B 96, 100403(R) (2017)
47. Hohlfeld, J., Matthias, E., Knorren, R., Bennemann, K.H.: Nonequilibrium magnetization
dynamics of nickel. Phys. Rev. Lett. 78, 4861–4864 (1997)
48. Miltat, J., Albuquerque, G., Thiaville, A.: An introduction to micromagnetics in the dynamic
regime. In: Hillebrands, B., Ounadjela, K. (eds.) Spindynamics in Confined Magnetic
Structures I, vol. 83, pp. 1–33. Springer (2002)
49. Hohlfeld, J., Gerrits, T., Bilderbeek, M., Rasing, T., Awano, H., Ohta, N.: Fast magnetization
reversal of GdFeCo induced by femtosecond laser pulses. Phys. Rev. B 65, 012413 (2001)
50. Ju, G., Hohlfeld, J., Bergman, B., van de Veerdonk, R.J.M., Mryasov, O.N., Kim, J.-Y., Wu,
X., Weller, D., Koopmans, B.: Ultrafast generation of ferromagnetic order via laser-induced
phase transformation in FeRh thin films. Phys. Rev. Lett. 93, 197403 (2004)
51. Thiele, J.-U., Buess, M., Back, C.H.: Spin dynamics of the ferromagnetic-to-
antiferromagnetic phase transition in FeRh on a sub-picosecond time scale. Appl. Phys. Lett.
85, 2857–2859 (2004)
52. Radu, I., Stamm, C., Pontius, N., Kachel, T., Ramm, P., Thiele, J.-U., Dürr, H.A., Back,
C.H.: Laser-induced generation and quenching of magnetization on FeRh studied with time-
resolved X-ray magnetic circular dichroism. Phys. Rev. B 81, 104415 (2010)
53. Pressacco, F., Uhlír̃, V., Gatti, M., Nicolaou, A., Bendounan, A., Arregi, J.A., Patel, S.K.K.,
Fullerton, E.E., Krizmancic, D., Sirotti, F.: Laser induced phase transition in epitaxial FeRh
layers studied by pump-probe valence band photoemission (2018). arXiv:1803.00780
54. van Kampen, M., Koopmans, B., Kohlhepp, J.T., de Jonge, W.J.M.: Laser-induced precession
in canted-spin ferromagnets. J. Magn. Magn. Mat. 240, 291–293 (2002)
55. Koopmans, B., Ruigrok, J.J.M., Longa, F.D., de Jonge, W.J.M.: Unifying ultrafast magneti-
zation dynamics. Phys. Rev. Lett. 95, 267207 (2005)
56. van Kampen, M., Jozsa, C., Kohlhepp, J.T., LeClair, P., Lagae, L., de Jonge, W.J.M.,
Koopmans, B.: All-optical probe of coherent spin waves. Phys. Rev. Lett. 88, 227201
(2002)
57. Ju, G., Nurmikko, A.V., Farrow, R.F.C., Marks, R.F., Carey, M.J., Gurney, B.A.: Ultrafast
optical modulation of an exchange biased ferromagnetic/antiferromagnetic bilayer. Phys. Rev.
B 58, R11857–R11860 (1998)
540 M. L. M. Lalieu and B. Koopmans

58. Hübner, W., Zhang, G.P.: Ultrafast spin dynamics in nickel. Phys. Rev. B 58, R5920–R5920
(1998)
59. Bigot, J.-Y., Vomir, M., Beaurepaire, E.: Coherent ultrafast magnetism induced by femtosec-
ond laser pulses. Nat. Phys. 5, 515–520 (2009)
60. Stupakiewicz, A., Szerenos, K., Afanasiev, D., Kirilyuk, A., Kimel, A.V.: Ultrafast nonther-
mal photo-magnetic recording in a transparent medium. Nature 542, 71–74 (2017)
61. van der Ziel, J.P., Pershan, P.S., Malmstrom, L.D.: Optically-induced magnetization resulting
from the inverse Faraday effect. Phys. Rev. Lett. 15, 190–193 (1965)
62. Agranat, M.B., Ashikov, S.I., Granovskii, A.B., Rukman, G.I.: Interaction of picosecond laser
pulses with the electron, spin and phonon subsystem of nickel. Zh. Eksp. Teor. Fiz. 86, 1376–
1379 (1984)
63. Vaterlaus, A., Beutler, T., Meier, F.: Spin-lattice relaxation time of ferromagnetic gadolinium
determined with time-resolved spin-polarized photoemission. Phys. Rev. Lett. 67, 3314–3317
(1991)
64. Vaterlaus, A., Beutler, T., Guarisco, D., Lutz, M., Meier, F.: Spin-lattice relaxation in
ferromagnets studied by time-resolved spin-polarized photoemission. Phys. Rev. B 46, 5280–
5286 (1992)
65. Hübner, W., Bennemann, K.H.: Simple theory for spin-lattice relaxation in metallic rare-earth
ferromagnets. Phys. Rev. B 53, 3422–3427 (1996)
66. Scholl, A., Baumgarten, L., Jacquemin, R., Eberhardt, W.: Ultrafast spin dynamics of
ferromagnetic thin films observed by fs spin-resolved two-photon photemission. Phys. Rev.
Lett. 79, 5146–5149 (1997)
67. Koopmans, B., van Kampen, M., Kohlhepp, J.T., de Jonge, W.J.M.: Ultrafast magneto-optics
in nickel: magnetism or optics? Phys. Rev. Lett. 85, 844–847 (2000)
68. Koopmans, B., van Kampen, M., de Jonge, W.J.M.: Experimental access to femtosecond spin
dynamics. J. Phys. Cond. Mat. 15, S723–S736 (2003)
69. Regensburger, H., Vollmer, R., Kirschner, J.: Time-resolved magnetization-induced
second-harmonic generation from the Ni(110) surface. Phys. Rev. B 61, 14716–14719
(2000)
70. Guidoni, L., Beaurepaire, E., Bigot, J.-Y.: Magneto-optics in the ultarfast regime: thermaliza-
tion of spin populations in ferromagnetic films. Phys. Rev. Lett. 89, 017401 (2002)
71. Oppeneer, P.M., Liebsch, A.: Ultrafast demagnetization in nickel: theory of magneto-optics
for non-equilibrium electron distributions. J. Phys. Cond. Mat. 16, 5519–5530 (2004)
72. Dewhurst, J.K., Shallcross, S., Gross, E.K.U., Sharma, S.: Substrate-controlled ultrafast spin
injection and demagnetization. Phys. Rev. A 10, 044065 (2018)
73. Dewhurst, J.K., Elliott, P., Shallcross, S., Gross, E.K.U., Sharma, S.: Laser-induced intersite
spin transfer. Nano Lett. 18, 1842–1848 (2018)
74. Beaurepaire, E., Maret, M., Halté, V., Merle, J.-C., Daunois, A., Bigot, J.-Y.: Spin dynamics
in CoPt3 alloy films: a magnetic phase transition in the femtosecond time scale. Phys. Rev. B
58, 12134–12137 (1998)
75. Wietstruk, M., Melnikov, A., Stamm, C., Kachel, T., Pontius, N., Sultan, M., Gahl, C.,
Weinelt, M., Dürr, H.A., Bovensiepen, U.: Hot-electron-driven enhancement of spin-lattice
coupling in Gd and Tb 4f ferromagnets observed by femtosecond X-ray magnetic circular
dichroism. Phys. Rev. Lett. 106, 127401 (2011)
76. Müller, G.M., Walowski, J., Djordjevic, M., Miao, G-X., Gupta, A., Ramos, A.V., Gehrke, K.,
Moshnyaga, V., Samwer, K., Schmalhorst, J., Thomas, A., Hütten, A., Reiss, G., Moodera,
J.S., Münzenberg, M.: Spin polarization in half-metals probed by femtosecond spin excitation.
Nat. Mater. 8, 56–61 (2009)
77. Ogasawara, T., Ohgushi, K., Tomioka, Y., Takahashi, K.S., Okamoto, H., Kawasaki, M.,
Tokura, Y.: General features of photoinduced spin dynamics in ferromagnetic and ferrimag-
netic compounds. Phys. Rev. Lett. 94, 087202 (2005)
78. Steil, D., Alebrand, S., Roth, T., Krauß, M., Kubota, T., Oogane, M., Ando, Y., Schneider,
H.C., Aeschlimann, M., Cinchetti, M.: Band-structure-dependent demagnetization in the
Heusler alloy Co2 Mn1−x Fex Si. Phys. Rev. Lett. 105, 217202 (2010)
10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films 541

79. Wanh, J., Khodaparast, G.A., Kono, J., Slupinski, T., Oiwa, A., Munekata, H.: Ultrafast
optical manipulation of ferromagentic order in InMnAs/GaSb. J. Supercond. 16, 373–377
(2003)
80. Wang, J., Sun, C., Kono, J., Oiwa, A., Munekata, H., Cywinski, L., Sham, J.J.: Ultrafast
quenching of ferromagnetism in InMnAs induced by intense laser irradiation. Phys. Rev. Lett.
95, 167401 (2005)
81. Kise, T., Ogasawara, T., Ashida, M., Tomioka, Y., Tokura, Y., Kuwata-Gonokami, M.:
Ultrafast spin dynamics and critical behavior in half-metallic ferromagnet: Sr2 FeMoO6 . Phys.
Rev. Lett. 85, 1986–1989 (2000)
82. Rhie, H.S., Dürr, H.A., Eberhardt, W.: Femtosecond electron and spin dynamics in Ni/W(110)
films. Phys. Rev. Lett. 90, 247201 (2003)
83. Dürr, H.A., Stamm, C., Kachel, T., Pontius, N., Mitzner, R., Quast, T., Holldack, K., Khan, S.,
Lupulescu, C., Aziz, E.F., Wietstruk, M., Eberhardt, W.: Ultrafast electron and spin dynamics
in nickel probed with femtosecond X-ray pulses. IEEE Trans. Magn. 44, 1957–1961 (2008)
84. Boeglin, C., Beaurepaire, E., Halté, V., López-Flores, V., Stamm, C., Pontius, N., Dürr, H.A.,
Bigot, J.-Y.: Distinguishing the ultrafast dynamics of spin and orbital moments in solids.
Nature 465, 458–461 (2010)
85. Roth, T., Schellekens, A.J., Alebrand, S., Schmitt, O., Steil, D., Koopmans, B., Cinchetti, M.,
Aeschlimann, M.: Temperature dependence of laser-induced demagnetization in Ni: a key for
identifying the underlying mechanism. Phys. Rev. X 2, 021006 (2012)
86. Bonetti, S., Hoffmann, M.C., Sher, M.-J., Chen, Z., Yang, S.-H., Samant, M.G., Parkin, S.S.P.,
Dürr, H.A.: THz-driven ultrafast spin-lattice scattering in amorphous metallic ferromagnets.
Phys. Rev. Lett. 117, 087205 (2016)
87. Beaurepaire, E., Turner, G.M., Harrel, S.M., Beard, M.C., Bigot, J.-Y., Schmuttenmaer, C.A.:
Coherent terahertz emission from ferromagnetic films excited by femtosecond laser pulses.
Appl. Phys. Lett. 84, 3465–3467 (2004)
88. Hilton, D.J., Averitt, R.D., Meserole, C.A., Fisher, G.L., Funk, D.J., Thompson, J.D., Taylor,
A.J.: Terahertz emission via ultrashort-pulse excitation of magnetic metal films. Opt. Lett. 29,
1805–1807 (2004)
89. Huisman, T.J., Mikhaylovskiy, R.V., Tsukamoto, A., Rasing, T., Kimel, A.V.: Simultaneous
measurements of terahertz emission and magneto-optical kerr effect for resolving ultrafast
laser-induced demagnetization dynamics. Phys. Rev. B 92, 104419 (2015)
90. Bergeard, N., Hehn, M., Mangin, S., Lengaigne, G., Montaigne, F., Lalieu, M.L.M.,
Koopmans, B., Malinowski, G.: Hot-electron-induced ultrafast demagnetization in Co/Pt
multilayers. Phys. Rev. Lett. 117, 147203 (2016)
91. Ferté, T., Bergeard, N., Malinowski, G., Abrudan, R., Kachel, T., Holldack, K., Hehn, M.,
Boeglin, C.: Ultrafast hot-electron induced quenching of Tb 4f magnetic order. Phys. Rev. B
96, 144427 (2017)
92. Wilson, R.B., Yang, Y., Gorchon, J., Lambert, C-H., Salahuddin, S., Bokor, J.: Electric current
induced ultrafast demagnetization. Phys. Rev. B 96, 045105 (2017)
93. Schellekens, A.J., Verhoeven, W., Vader, T.N., Koopmans, B.: Investigating the contribution
of superdiffusive transport to ultrafast demagnetization of ferromagnetic thin films. Appl.
Phys. Lett. 102, 252408 (2013)
94. Einstein, A., de Haas, W.J.: Experimental proof of the existence of Ampére’s molecular
currents. KNAW Proc. 18 I, 696–711 (1915)
95. Frenkel, V.Y.: On the history of the Einstein’de Haas effect. Usp. Fiz. Nauk 128, 545–557
(1979)
96. Töws, W., Pastor, G.M.: Many-body theory of ultrafast demagnetization and angular momen-
tum transfer in ferromagnetic transition metals. Phys. Rev. Lett. 115, 217204 (2015)
97. Krauß, M., Roth, T., Alebrand, S., Steil, D., Cinchetti, M., Aeschlimann, M., Schneider,
H.C.: Ultrafast demagnetization of ferromagnetic transition metals: the role of the coulomb
interaction. Phys. Rev. B 80, 180407 (2009)
98. Fähnle, M., Haag, M., Illg, C.: Is the angular momentum of a ferromagnetic sample after
exposure to a fs laser pulse conserved? J. Magn. Magn. Mat. 347, 45–46 (2013)
542 M. L. M. Lalieu and B. Koopmans

99. Elliott, R.J.: Theory of the effect of spin-orbit coupling on magnetic resonance in some
semiconductors. Phys. Rev. 96, 266–279 (1954)
100. Rhie, H.S., Dürr, H.A., Eberhardt, W.: Femtosecond evolution of electronic interactions at the
Ni(111) surface. Appl. Phys. A 82, 9–14 (2005)
101. Aeschlimann, M., Bauer, M., Pawlik, S., Weber, W., Burgermeister, R., Oberli, D., Siegmann,
H.C.: Ultrafast spin-dependent electron dynamics in fcc Co. Phys. Rev. Lett. 79, 5158–5161
(1997)
102. Dalla Longa, F., Kohlhepp, J.T., de Jonge, W.J.M., Koopmans, B.: Influence of photon angular
momentum on ultrafast demagnetization in nickel. Phys. Rev. B 75, 224431 (2007)
103. Koopmans, B., Malinowski, G., Dalla Longa, F., Steiauf, D., Fahnle, M., Roth, T., Cinchetti,
M., Aeschlimann, M.: Explaining the paradoxical diversity of ultrafast laser-induced demag-
netization. Nat. Mater. 9, 259–265 (2010)
104. Zhang, G.P., Hübner, W.: Femtosecond spin dynamics in the time domain. J. Appl. Phys. 85,
5657–5659 (1999)
105. Zhang, G.P., Hübner, W.: Laser-induced ultrafast demagnetization in ferromagnetic metals.
Phys. Rev. Lett. 85, 3025–3028 (2000)
106. Zhang, G.P., Hübner, W., Lefkidis, G., Bai, Y., George, T.F.: Paradigm of the time-resolved
magneto-optical Kerr effect for femtosecond magnetism. Nat. Phys. 5, 499–502 (2009)
107. Carva, K., Battiato, M., Oppeneer, P.M.: Is the controversy over femtosecond magneto-optics
really solved? Nat. Phys. 7, 665 (2011)
108. Atxitia, U., Chubykalo-Fesenko, O.: Ultrafast magnetization dynamics rates within the
Landau-Lifshitz-Bloch model. Phys. Rev. B 84, 144414 (2011)
109. Kazantseva, N., Nowak, U., Chantrell, R.W., Hohlfeld, J., Rebei, A.: Slow recovery of the
magnetisation after a sub-picosecond heat pulse. Europhys. Lett. 81, 27004 (2008)
110. Evans, R.F.L., Fan, W.J., Chureemart, P., Ostler, T.A., Ellis, M.O.A., and Chantrell, R.W.:
Atomistic spin model simulations of magnetic nanomaterials. J. Phys. Condens. Matter. 26,
103202 (2014)
111. Steiauf, D., Fähnle, M.: Elliott-Yafet mechanism and the discussion of femtosecond magne-
tization dynamics. Phys. Rev. B 79, 140401 (2009)
112. Schellekens, A.J., Koopmans, B.: Comparing ultrafast demagnetization rates between com-
peting models for finite temperature magnetism. Phys. Rev. Lett. 110, 217204 (2013)
113. Tveten, E.G., Brataas, A., Tserkovnyak, Y.: Electron-magnon scattering in magnetic het-
erostructures far out of equilibrium. Phys. Rev. B 92, 180412(R) (2015)
114. Mueller, B.Y., Baral, A., Vollmar, S., Cinchetti, M., Aeschlimann, M., Schneider, H.C.,
Rethfeld, B.: Feedback effect during ultrafast demagnetization dynamics in ferromagnets.
Phys. Rev. Lett. 111, 167204 (2013)
115. Mueller, B.Y., Rethfeld, B.: Thermodynamic μT model of ultrafast magnetization dynamics.
Phys. Rev. B 90, 144420 (2014)
116. Krieger, K., Dewhurst, J.K., Elliott, P., Sharma, S., Gross, E.K.U.: Laser-induced demag-
netization at ultrashort time scales: predictions of TDDFT. J. Chem. Theory Comput. 11,
4870–4874 (2015)
117. Elliott, P., Müller, T., Dewhurst, J.K., Sharma, S., Gross, E.K.U.: Ultrafast laser induced local
magnetization dynamics in Heusler compounds. Scient. Rep. 6, 38911 (2016)
118. Fognini, A., Salvatella, G., Gort, R., Michlmayr, T., Vaterlaus, A., Acremann, Y.: The
influence of the excitation pulse length on ultrafast magnetization dynamics in nickel. Struct.
Dyn. 2, 024501 (2015)
119. Carva, K., Battiato, M., Legut, D., Oppeneer, P.M.: Ab initio theory of electron-phonon medi-
ated ultrafast spin relaxation of laser-excited hot electrons in transition-metal ferromagnets.
Phys. Rev. B 87, 184425 (2013)
120. Kuiper, K.C., Roth, T., Schellekens, A.J., Schmitt, O., Koopmans, B., Cinchetti, M., Aeschli-
mann, M.: Spin-orbit enhanced demagnetization rate in Co/Pt-multilayers. Appl. Phys. Lett.
105, 202402 (2014)
121. Melnikov, A., Radu, I., Bovensiepen, U., Krupin, O., Starke, K., Matthias, E., Wolf, M.:
Coherent optical phonons and parametrically coupled magnons induced by femtosecond laser
excitation of the Gd(0001) surface. Phys. Rev. Lett. 91, 227403 (2003)
10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films 543

122. Sultan, M., Atxitia, U., Melnikov, A., Chubykalo-Fesenko, O., Bovensiepen, U.: Electron-
and phonon-mediated ultrafast magnetization dynamics of Gd(0001). Phys. Rev. B 85,
184407 (2012)
123. Frietsch, B., Carley, R., Gleich, M., Teichmann, M., Bowlan, J., Weinelt, M.: Fluence-
dependent dynamics of the 5d6s exchange splitting in Gd metal after femtosecond laser
excitation. Jpn. J. Appl. Phys. 55, 07MD02 (2016)
124. Radu, I., Woltersdorf, G., Kiessling, M., Melnikov, A., Bovensiepen, U., Thiele, J.-U., Back,
C.H.: Laser-induced magnetization dynamics of lanthanide-doped permalloy thin films. Phys.
Rev. Lett. 102, 117201 (2009)
125. Turgut, E., Zusin, D., Legut, D., Carva, K., Knut, R., Shaw, J.M., Chen, C., Tao, Z., Nembach,
H.T., Silva, T.J., Mathias, S., Aeschlimann, M., Oppeneer, P.M., Kapteyn, H.C., Murnane,
M.M., Grychtol, P.: Stoner versus Heisenberg: ultrafast exchange reduction and magnon
generation during laser-induced demagnetization. Phys. Rev. B 94, 220408(R) (2016)
126. Eich, S., Plötzing, M., Rollinger, M., Emmerich, S., Adam, R., Chen, C., Kapteyn, H.C.,
Murnane, M.M., Plucinski, L., Steil, D., Stadtmüller, B., Cinchetti, M., Aeschlimann, M.,
Schneider, C.M., Mathias, S.: Band structure evolution during the ultrafast ferromagnetic-
paramagnetic phase transition in cobalt. Sci. Adv. 3, e1602094 (2017)
127. Gort, R., Bühlmann, K., Däster, S., Salvatella, G., Hartmann, N., Zemp, Y., Holenstein, S.,
Stieger, C., Fognini, A., Michlmayr, T.U., Bähler, T., Vaterlaus, A., Acremann, Y.: Early
stages of ultrafast spin dynamics in a 3d ferromagnet. Phys. Rev. Lett. 121, 087206 (2018)
128. Pershan, P.S., van der Ziel, J.P., Malmstrom, L.D.: Theoretical discussion of the inverse
Faraday effect, Raman scattering, and related phenomena. Phys. Rev. 143, 574–583
(1966)
129. Kimel, A.V., Kirilyuk, A., Usachev, P.A., Pisarev, R.V., Balbashov, A.M., Rasing, T.: Ultrafast
non-thermal control of magnetization by instantaneous photomagnetic pulses. Nature 435,
655–657 (2005)
130. Rebei, A., Hohfeld, J.: The magneto-optical barnett effect: Circularly polarized light induced
femtosecond magnetization reversal. Phys. Lett. A 372, 1915–1918 (2008)
131. Hohlfeld, J., Stanciu, C.D., Rebei, A.: Athermal all-optical femtosecond magnetization
reversal in GdFeCo. Appl. Phys. Lett. 94, 152504 (2009)
132. Vahaplar, K., Kalashnikova, A.M., Kimel, A.V., Hinzke, D., Nowak, U., Chantrell, R.,
Tsukamoto, A., Itoh, A., Kirilyuk, A., Rasing, T.: Ultrafast path for optical magnetization
reversal via a strongly nonequilibrium state. Phys. Rev. Lett. 103, 117201 (2009)
133. Steil, D., Alebrand, S., Hassdenteufel, A., Cinchetti, M., Aeschlimann, M.: All-optical
magnetization recording by tailoring optical excitation parameters. Phys. Rev. B 84, 224408
(2011)
134. Vahaplar, K., Kalashnikova, A.M., Kimel, A.V., Gerlach, S., Hinzke, D., Nowak, U.,
Chantrell, R., Tsukamoto, A., Itoh, A., Kirilyuk, A, Rasing, T.: All-optical magnetization
reversal by circularly polarized laser pulses: Experiment and multiscale modeling. Phys. Rev.
B 85, 104402 (2012)
135. Alebrand, S., Hassdenteufel, A., Steil, D., Cinchetti, M., Aeschlimann, M.: Interplay of
heating and helicity in all-optical magnetization switching. Phys. Rev. B 85, 092401 (2012)
136. Mentink, J.H., Hellsvik, J., Afanasiev, D.V., Ivanov, B.A., Kirilyuk, A., Kimel, A.V., Eriksson,
O., Katsnelson, M.I., Rasing, T.: Ultrafast spin dynamics in multisublattice magnets. Phys.
Rev. Lett. 108, 057202 (2012)
137. Khorsand, A.R., Savoini, M., Kirilyuk, A., Kimel, A.V., Tsukamoto, A., Itoh, A., Rasing, T.:
Role of magnetic circular dichroism in all-optical magnetic recording. Phys. Rev. Lett. 108,
127205 (2012)
138. Wienholdt, S., Hinzke, D., Carva, K., Oppeneer, P.M., Nowak, U.: Orbital-resolved spin
model for thermal magnetization switching in rare-earth-based ferrimagnets. Phys. Rev. B
88, 020406 (2013)
139. Evans, R.F.L., Ostler, T.A., Chantrell, R.W., Radu, I., Rasing, T.: Ultrafast thermally induced
magnetic switching in synthetic ferrimagnets. Appl. Phys. Lett. 104, 082410 (2014)
140. Schellekens, A.J., Koopmans, B.: Microscopic model for ultrafast magnetization dynamics of
multisublattice magnets. Phys. Rev. B 87, 020407 (2013)
544 M. L. M. Lalieu and B. Koopmans

141. Alebrand, S., Gottwald, M., Hehn, M., Steil, D., Cinchetti, M., Lacour, D., Fullerton, E.E.,
Aeschlimann, M., Mangin, S.: Light-induced magnetization reversal of high-anisotropy TbCo
alloy films. Appl. Phys. Lett. 101, 162408 (2012)
142. Hassdenteufel, A., Schubert, C., Schmidt, J., Richter, P., Zahn, D.R.T., Salvan, G., Helm, M.,
Bratschitsch, R., Albrecht, M.: Dependence of all-optical magnetic switching on the sublattice
magnetization orientation in Tb-Fe thin films. Appl. Phys. Lett. 105, 112403 (2014)
143. Cornelissen, T.D., Córdoba, R., Koopmans, B.: Microscopic model for all optical switching
in ferromagnets. Appl. Phys. Lett. 108, 142405 (2016)
144. Berritta, M., Mondal, R., Carva, K., Oppeneer, P.M.: Ab initio theory of coherent laser-
induced magnetization in metals. Phys. Rev. Lett. 117, 137203 (2016)
145. El Hadri, M.S., Pirro, P., Lambert, C.-H., Petit-Watelot, S., Quessab, Y., Hehn, M., Montaigne,
F., Malinowski, G., Mangin, S.: Two types of all-optical magnetization switching mechanisms
using femtosecond laser pulses. Phys. Rev. B 94, 064412 (2016)
146. Medapalli, R., Afanasiev, D., Kim, D.K., Quessab, Y., Manna, S., Montoya, S.A., Kirilyuk,
A., Rasing, T., Kimel, A.V., Fullerton, E.E.: Multiscale dynamics of helicity-dependent all-
optical magnetization reversal in ferromagnetic co/pt multilayers. Phys. Rev. B 96, 224421
(2017)
147. Quessab, Y., Medapalli, R., El Hadri, M.S., Hehn, M., Malinowski, G., Fullerton, E.E.,
Mangin, S.: Helicity-dependent all-optical domain wall motion in ferromagnetic thin films.
Phys. Rev. B 97, 054419 (2018)
148. El Hadri, M.S., Hehn, M., Pirro, P., Lambert, C-H., Malinowski, G., Fullerton, E.E., Mangin,
S.: Domain size criterion for the observation of all-optical helicity-dependent switching in
magnetic thin films. Phys. Rev. B 94, 064419 (2016)
149. Hassdenteufel, A., Schmidt, J., Schubert, C., Hebler, B., Helm, M., Albrecht, M., Brats-
chitsch, R.: Low-remanence criterion for helicity-dependent all-optical magnetic switching
in ferrimagnets. Phys. Rev. B 91, 104431 (2015)
150. Takahashi, Y.K., Medapalli, R., Kasai, S., Wang, J., Ishioka, K., Wee, S.H., Hellwig, O.,
Hono, K., Fullerton, E.E.: Accumulative magnetic switching of ultrahigh-density recording
media by circularly polarized light. Phys. Rev. A 6, 054004 (2016)
151. Gorchon, J., Yang, Y., Bokor, J.: Model for multishot all-thermal all-optical switching in
ferromagnets. Phys. Rev. B 94, 020409 (2016)
152. Ellis, M.O.A., Fullerton, E.E., Chantrell, R.W.: All-optical switching in granular ferromagnets
caused by magnetic circular dichroism. Sci. Rep. 6, 30522 (2016)
153. John, R., Berritta, M., Hinzke, D., Müller, C., Santos, T., Ulrichs, H., Nieves, P., Walowski, J.,
Mondal, R., Chubykalo-Fesenko, O., McCord, J., Oppeneer, P.M., Nowak, U., Münzenberg,
M.: Magnetisation switching of FePt nanoparticlerecording medium by femtosecond laser
pulses. Scient. Rep. 7, 4114 (2017)
154. Wilson, R.B., Gorchon, J., Yang, Y., Lambert, C-H., Salahuddin, S., Bokor, J.: Ultrafast
magnetic switching of gdfeco with electronic heat currents. Phys. Rev. B 95, 180409 (2017)
155. Xu, Y., Deb, M., Malinowski, G., Hehn, M., Zhao, W., Mangin, S.: Ultrafast magnetization
manipulation using single femtosecond light and hot-electron pulses. Adv. Mater. 29,
1703474 (2017)
156. Yang, Y., Wilson, R.B., Gorchon, J., Lambert, C-H., Salahuddin, S., Bokor, J.: Ultrafast
magnetization reversal by picosecond electrical pulses. Sci. Adv. 3, e1603117 (2017)
157. Gorchon, J., Lambert, C-H., Yang, Y., Pattabi, A., Wilson, R.B., Salahuddin, S., Bokor, J.:
Single shot ultrafast all optical magnetization switching of ferromagnetic Co/Pt multilayers.
Appl. Phys. Lett. 111, 042401 (2017)
158. Iihama, S., Xu, Y., Deb, M., Malinowski, G., Hehn, M., Gorchon, J., Fullerton, E.E., Mangin,
S.: Single-shot multi-level all-optical magnetization switching mediated by spin transport.
Adv. Mater. 30, 1804004 (2018)
159. Choi, G-M., Min, B-C.: Laser-driven spin generation in the conduction bands of ferrimagnetic
metals. Phys. Rev. B 97, 014410 (2018)
160. Lalieu, M.L.M., Peeters, M.J.G., Haenen, S.R.R., Lavrijsen, R., Koopmans, B.: Deterministic
all-optical switching of synthetic ferrimagnets using single femtosecond laser pulses. Phys.
Rev. B 96, 220411(R) (2017)
10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films 545

161. Lalieu, M.L.M., Lavrijsen, R., Koopmans, B.: Integrating all-optical switching with spintron-
ics. Nat. Commun. 10, 110 (2019)
162. Gerlach, S., Oroszlany, L., Hinzke, D., Sievering, S., Wienholdt, S., Szunyogh, L., Nowak, U.:
Modeling ultrafast all-optical switching in synthetic ferrimagnets. Phys. Rev. B 95, 224435
(2017)
163. Parkin, S., Yang, S-H.: Memory on the racetrack. Nat. Nanotechnol. 10, 195–198 (2015)
164. Yang, S-H., Ryu, K-S., Parkin, S.: Domain-wall velocities of up to 750 m s−1 driven by
exchange-coupling torque in synthetic antiferromagnets. Nat. Nanotechnol. 10, 221–226
(2015)
165. Kim, K-J., Kim, S.K., Hirata, Y., Oh, S-H., Tono, T., Kim, D-H., Okuno, T., Ham, W.S.,
Kim, S., Go, G., Tserkovnyak, Y., Tsukamoto, A., Moriyama, T., Lee, K-J., Ono, T.: Fast
domain wall motion in the vicinity of the angular momentum compensation temperature of
ferrimagnets. Nat. Mater. 16, 1187–1192 (2017)
166. Caretta, L., Mann, M., Büttner, F., Ueda, K., Pfau, B., C.M. Günther, Hessing, P., Churikova,
A., Klose, C., Schneider, M., Engel, D., Marcus, C., Bono, D., Bagschik, K., Eisebitt, S.,
Beach, G.S.D.: Fast current-driven domain walls and small skyrmions in a compensated
ferrimagnet. Nat. Nanotechnol. 13, 1154–1160 (2018)
167. Kent, A.D., Worledge, D.C.: A new spin on magnetic memories. Nat. Nanotechnol. 10, 187–
191 (2015)
168. Melnikov, A., Razdolski, I., Wehling, T.O., Papaioannou, E.T., Roddatis, V., Fumagalli, P.,
Aktsipetrov, O., Lichtenstein, A.I., Bovensiepen, U.: Ultrafast transport of laser-excited spin-
polarized carriers in Au/Fe/MgO(001). Phys. Rev. Lett. 107, 076601 (2011)
169. Battiato, M., Carva, K., Oppeneer, P.M.: Theory of laser-induced ultrafast superdiffusive spin
transport in layered heterostructures. Phys. Rev. B 86, 024404 (2012)
170. Rudolf, D., La-O-Vorakiat, C., Battiato, M., Adam, R., Shaw, J.M., Turgut, E., Maldonado,
P., Mathias, S., Grychtol, P., Nembach, H.T., Silva, T.J., Aeschlimann, M., Kapteyn, H.C.,
Murnane, M.M., Schneider, C.M., Oppeneer, P.M.: Ultrafast magnetization enhancement in
metallic multilayers driven by superdiffusive spin current. Nat. Commun. 3, 1037 (2012)
171. Eschenlohr, A., Battiato, M., Maldonado, P., Pontius, N., Kachel, T., Holldack, K., Mitzner,
R., Föhlisch, A., Oppeneer, P.M., Stamm, C.: Ultrafast spin transport as key to femtosecond
demagnetization. Nat. Mater. 12, 332–336 (2013)
172. Khorsand, A.R., Savoini, M., Kirilyuk, A., Rasing, T.: Optical excitation of thin magnetic
layers in multilayer structures. Nat. Mater. 13, 101–102 (2014)
173. Eschenlohr, A., Battiato, M., Maldonado, P., Pontius, N., Kachel, T., Holldack, K., Mitzner,
R., Föhlisch, A., Oppeneer, P.M., Stamm, C.: Optical excitation of thin magnetic layers in
multilayer structures Reply. Nat. Mater. 13, 102–103 (2014)
174. Eschenlohr, A., Persichetti, L., Kachel, T., Gabureac, M., Gambardella, P., Stamm, C.: Spin
currents during ultrafast demagnetization of ferromagnetic bilayers. J. Phys. Condens. Matter.
29, 384002 (2017)
175. Choi, G-M., Cahill, D.G.: Kerr rotation in Cu, Ag, and Au driven by spin accumulation and
spin-orbit coupling. Phys. Rev. B 90, 214432 (2014)
176. Choi, G-M., Moon, C-H., Min, B-C., Lee, K-J., Cahill, D.G.: Thermal spin-transfer torque
driven by the spin-dependent Seebeck effect in metallic spin valves. Nat. Phys. 11, 576–582
(2015)
177. Alekhin, A., Razdolski, I., Ilin, N., Meyburg, J.P., Diesing, D., Roddatis, V., Rungger, I.,
Stamenova, M., Sanvito, S., Bovensiepen, U., Melnikov, A.: Femtosecond spin current pulses
generated by the nonthermal spin-dependent Seebeck effect and interacting with ferromagnets
in spin valves. Phys. Rev. Lett. 119, 017202 (2017)
178. Razdolski, I., Alekhin, A., Ilin, N., Meyburg, J.P., Roddatis, V., Diesing, D., Bovensiepen,
U., Melnikov, A.: Nanoscale interface confinement of ultrafast spin transfer torque driving
non-uniform spin dynamics. Nat. Commun. 8, 15007 (2017)
179. Lalieu, M.L.M., Helgers, P.L.J., Koopmans, B.: Absorption and generation of femtosecond
laser-pulse excited spin currents in noncollinear magnetic bilayers. Phys. Rev. B 96, 014417
(2017)
180. Stiles, M.D., Zangwill, A.: Anatomy of spin-transfer torque. Phys. Rev. B 66, 014407 (2002)
546 M. L. M. Lalieu and B. Koopmans

181. Baláž, P., Žonda, M., Carva, K., Maldonado, P., Oppeneer, P.M.: Transport theory for
femtosecond laser-induced spin-transfer torques. J. Phys. Condens. Matter 30, 115801 (2018)
182. Alekhin, A., Bürstel, D., Melnikov, A., Diesing, D., Bovensiepen, U.: Ultrafast laser-excited
spin transport in Au/Fe/MgO(001), relevance of the Fe layer thickness. In: Bigot, J.Y., Hubner,
W., Rasing, T., Chantrell, R. (eds.) Ultrafast Magnetism I, vol. 159, pp. 241–243. Springer
Proceedings in Physics (2014)
183. Ulrichs, H., Razdolski, I.: Micromagnetic view on ultrafast magnon generation by femtosec-
ond spin current pulses. Phys. Rev. B 98, 054429 (2018)
184. Hellman, F., Hoffmann, A., Tserkovnyak, Y., Beach, G.S.D., Fullerton, E.E., Leighton, C.,
MacDonald, A.H., Ralph, D.C., Arena, D.A., Dürr, H.A., Fischer, P., Grollier, J., Heremans,
J.P., Jungwirth, T., Kimel, A.V., Koopmans, B., Krivorotov, I.N., May, S.J., Petford-Long,
A.K., Rondinelli, J.M., Samarth, N., Schuller, I.K., Slavin, A.N., Stiles, M.D., Tchernyshyov,
O., Thiaville, A., Zink, B.L.: Interface-induced phenomena in magnetism. Rev. Mod. Phys.
89, 025006 (2017)
185. Kirilyuk, A., Kimel, A.V., Rasing, T.: Ultrafast optical manipulation of magnetic order. Rev.
Mod. Phys. 82, 2731 (2010)
186. Koopmans, B.: Time-resolved Kerr-effect and spin dynamics in itinerant ferromagnets. In:
Kronmüller, H., Parkin, S. (eds.) Handbook of Magnetism and Advanced Magnetic Materials:
Volume 3 Novel Techniques for Characterizing and Preparing Samples, pp. 1589–1613. Wiley
(2007)

Further Reading
1. Qiu, Z.Q., Bader, S.D.: Surface magneto-optic kerr effect. [37]
2. Soldatov, I.V., Schäfer, R.: Advances in quantitative Kerr microscopy. Phys. Rev. B 95, 014426 (2017)
3. Oppeneer, P.M.: Magneto-optical Kerr spectra. [17]
4. Hellman, F., et al.: Chapter IV in “Interface-induced phenomena in magnetism”. [184]
5. Kirilyuk, A., et al.: Ultrafast optical manipulation of magnetic order. [185]
6. Koopmans, B.: Laser-induced magnetization dynamics. [26]
7. Koopmans, B.: Time-resolved Kerr-effect and Spin Dynamics in Itinerant Ferromagnets. [186]

Mark L. M. Lalieu received his PhD from the Eindhoven


University of Technology in 2019. Within the Physics of Nanos-
tructures group, he worked in the field of laser-induced ultrafast
magnetization dynamics and its integration with the fields of
spintronics and magnonics.
10 Magneto-optics and Laser-Induced Dynamics of Metallic Thin Films 547

Bert Koopmans received his PhD from the University of Gronin-


gen. After a short stay at the Radboud University Nijmegen, and
three years as a Humboldt Fellow at the Max-Planck Institute
in Stuttgart, he joined the Eindhoven University of Technology,
where since 2003 he chairs the group Physics of Nanostructures.
His research interests encompass spintronics, nanomagnetism and
ultrafast magnetization dynamics.
Magnetostriction and Magnetoelasticity
11
Dirk Sander

Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 550
Classification of Magnetoelastic Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 551
Anisotropic Magnetostriction (Joule Magnetostriction) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 551
Magnetovolume Effects, Spontaneous Magnetostriction, Forced
Magnetostriction, and Invar Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 552
Villari Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 553
E Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 553
Magnetomechanical Damping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 555
Wiedemann Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 556
Matteucci Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 557
Magnetic Field-Induced Strain Phenomena, Which Differ from Joule
Magnetostriction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 557
Magnetoelasticity and Joule Magnetostriction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 560
Derivation of the Magnetostrictive Strain Tensor: Cubic Case . . . . . . . . . . . . . . . . . . . . . . . 560
Magnetostriction of Polycrystalline Cubic Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 564
Derivation of the Magnetostrictive Strain Tensor: Hexagonal Case . . . . . . . . . . . . . . . . . . . 565
Magnetostriction of Polycrystalline Hexagonal Materials . . . . . . . . . . . . . . . . . . . . . . . . . . 566
Magnetostriction and Stress-Induced Magnetic Anisotropy . . . . . . . . . . . . . . . . . . . . . . . . . . . 567
Magnetoelastic Effects in Films . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 569
Experimental Determination of Magnetostriction and Magnetoelastic Coupling . . . . . . . . . . 570
Magnetoelastic Coupling in Films . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 571
Magnetostriction and Magnetoelasticity: Physical Origin and Insights from Theory . . . . . . . 574
Compilation of Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 577
Magnetoelastic and Elasticity Data for Bulk Transition Metals . . . . . . . . . . . . . . . . . . . . . . 577
Theoretical and Experimental Values of Magnetoelastic Coupling
Coefficients and Their Strain Dependence (Table 4) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 578
Magnetostriction Data of Amorphous Fe Alloys (Table 5) . . . . . . . . . . . . . . . . . . . . . . . . . . 579

D. Sander ()
Max Planck Institute of Microstructure Physics, Halle, Germany
e-mail: sander@mpi-halle.de

© Springer Nature Switzerland AG 2021 549


J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_11
550 D. Sander

Magnetostriction Data of Fe-Ga (Galfenol), Fe-Ge, FeAl, Fe-Si,


Fe-Ga-Al, and Fe-Ga-Ge Alloys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 579
Zero Magnetostriction Alloys and Soft Magnetic Materials . . . . . . . . . . . . . . . . . . . . . . . . . 580
Magnetostriction Data for Paramagnetic Metals and Alloys . . . . . . . . . . . . . . . . . . . . . . . . 582
Magnetostriction Data for Bi . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 582
Magnetostriction Data for Tb, Dy, and Ho . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 582
Magnetostriction Data of TbFe2 (Terfenol) and Tb27 Dy73 Fe2
(Terfenol-D) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 584
Magnetostriction Data of Oxide Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 585
Notations for Lattice Strain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 585
Relation Between λ and B for the Hexagonal System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 586
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 586

Abstract

The physical concepts magnetostriction and magnetoelasticity are presented.


Spontaneous volume magnetostriction and saturation linear magnetostriction
are distinguished. Various magnetoelastic phenomena are introduced, but the
emphasis is on magnetostriction in bulk samples and thin films. The equations
for magnetostrictive and magnetoelastic coefficients are derived for cubic,
hexagonal, and isotropic systems. Experiments on the measurement of the linear
magnetostriction λi and magnetoelastic coupling coefficients Bj are discussed.
Ab initio-based theory elucidates the physical origin of magnetostrictive effects
in metals at the electronic level, although accurate calculations are often elusive.
The magnetoelastic properties of nm thin films may deviate in magnitude and
sign from the bulk values. Both experiment and theory identify substrate-induced
lattice strain as a driving force for this deviation. Data on magnetostriction
and magnetoelasticity are compiled, including those of highly magnetostrictive
systems, such as (Tb,Dy)Fe2 (Terfenol) and (Fe,Ga) (Galfenol).

Introduction

Magnetostriction is the relative change of length of a sample when it is magnetized


[1,2,3,4,5,6,7,8,9,10,11,12,13]. It is customary to distinguish between spontaneous
volume magnetostriction, which is a change of volume in the magnetically ordered
state that is almost independent of applied magnetic field, and anisotropic linear
magnetostriction, also known as Joule magnetostriction, which saturates with the
magnetization.
The latter phenomenon was discovered by James Prescott Joule in 1841 [14] and
described in 1847 [15, 16]. Joule used an ingenious mechanical lever mechanism to
magnify 3000 times the tiny elongation upon magnetization of a 2-foot-long Fe bar
to make it measurable. He found a relative change of length upon magnetization of
order 1.4 × 10−6 . He also found that the net sample volume remained unchanged by
immersing the bar in water in a sealed container and looking for a displacement of
water in a capillary tube when the bar was magnetized. These historic observations
are consistent with today’s understanding of the phenomenon now known as Joule
magnetostriction.
11 Magnetostriction and Magnetoelasticity 551

Magnetostriction is not necessarily tiny. Although a typical order of magnitude is


10−5 for bulk 3d metals [17], values as large as 10−3 have been reported for highly
magnetostrictive alloys, such as FeTb (Terfenol) and GaFe (Galfenol), and for some
rare earth elements [18, 19, 20, 21, 22].
Anisotropic magnetostriction, or Joule magnetostriction, is the main focus of this
chapter. There are numerous other effects related to magnetoelasticity. These effects
describe the coupling between magnetization and elasticity, thermal expansion,
sample volume, strain, and torsion. They relate to the observed deformation of
a sample upon magnetization. The inverse effect, the change of the magnetic
configuration by imposing stress or torque on a sample, is also exploited to
study magnetostriction. These effects are described in section “Classification of
Magnetoelastic Effects”.
The derivation of the respective magnetostrictive coefficients from a measure-
ment of change of length upon magnetization is outlined in section “Magnetoe-
lasticity and Joule Magnetostriction”. It is shown that the underlying coupling
between lattice strain and magnetization direction is well described in the frame-
work of magnetoelasticity. The mathematical description of magnetoelasticity in
view of the tensor properties of crystal elasticity is elucidated, and corresponding
expressions for magnetostrictive coefficients and magnetoelastic coupling coef-
ficients are presented. Sections “Magnetostriction and Stress-Induced Magnetic
Anisotropy” and “Magnetoelastic Effects in Films” discuss stress-induced magnetic
anisotropy and magnetoelastic effects in thin films. Section “Experimental Deter-
mination of Magnetostriction and Magnetoelastic Coupling” describes experiments
to measure magnetostriction and magnetoelastic coupling in bulk and in films.
Insights from theory on the physical origin of magnetoelasticity on the electronic
level of metals and 3d oxides are presented in section “Magnetostriction and
Magnetoelasticity: Physical Origin and Insights from Theory”. A compilation
of magnetostrictive and magnetoelastic coupling data concludes this chapter in
section “Compilation of Data”.

Classification of Magnetoelastic Effects

Magnetostriction describes the strain of a sample upon magnetization. A further,


closely related manifestation of magnetoelasticity is the inverse effect: the impact
of an externally applied strain, e.g., via stress or torsion, on the magnetism of the
sample. This gives rise to numerous magnetoelastic effects, which are categorized
in this section.

Anisotropic Magnetostriction (Joule Magnetostriction)

Experiments show that an unmagnetized ferromagnet strains when it is magnetized


to saturation. It is observed that the strain along the magnetization direction differs
in sign and magnitude from the strain measured perpendicular to the magnetization
552 D. Sander

λ >0

M = Ms
H
λ <0
T < TC
ω>0
T > TC Mavg = 0
H=0

Fig. 1 Schematic view of magnetostriction in a ferromagnet. Below the Curie temperature TC , a


spontaneous lattice strain, the spontaneous volume magnetostriction ω, is observed, even in the
absence of a magnetic field. The application of a field H induces an anisotropic lattice strain.
Positive (negative) magnetostriction, λ > 0 ( λ < 0 ), gives an elongation (compression) along the
field direction and an opposite strain perpendicular to it. The sketch highly exaggerates the volume
and linear strains of order 10−6 to 10−2

direction so that the net volume is conserved. This situation is schematically


illustrated in Fig. 1.
Formally, this magnetization-induced lattice strain is ascribed to magnetoelastic
contributions to the energy density. Qualitatively, these contributions are expressed
by terms such as fme = B. Here, B is the magnetoelastic coupling coefficient
(units J/m3 ), and the dimensionless lattice strain is . The derivative of the
magnetoelastic energy density with respect to lattice strain gives a stress ∂fme /∂ =
τme = B, which strains the lattice to the point where elastic stresses compensate the
magnetization-induced stress [2, 17]. In general, a three-dimensional magnetostric-
tive strain state results. This qualitative description already points at the importance
of the interplay between magnetoelasticity and crystal elasticity for the magnitude
of magnetostriction. The link between the two becomes obvious in the quantitative
derivation of magnetostrictive strains below in section “Magnetoelasticity and Joule
Magnetostriction”.

Magnetovolume Effects, Spontaneous Magnetostriction, Forced


Magnetostriction, and Invar Effects

The spontaneous volume magnetostriction in Fig. 1 is defined as the difference in


atomic volume between the magnetically ordered state and the paramagnetic state.
Thus, the temperature-driven transition at the Curie temperature brings with it a
change of lattice volume, which deviates from the volume extrapolated using the
thermal expansion α(T ) observed above the Curie temperature. Consequently, the
11 Magnetostriction and Magnetoelasticity 553

temperature dependence α(T ) shows distinct variations near the magnetic ordering
temperature [4, 23]. This is also true for antiferromagnets and other types of
magnetic order.
The spontaneous volume magnetostriction of a metal can be slightly altered
by applying a large external magnetic field, and this gives rise to forced volume
magnetostriction [24].
The spontaneous volume magnetostriction, if positive, can produce a lattice
expansion to compensate or cancel the usual contraction of lattice spacing with
deceasing temperature, leading to vanishingly small or even negative thermal
expansion – lattice expansion upon cooling. The former effect is found in invar
alloys such as Fe64 Ni36 [25] and the latter in Mn3 (Cu(1−x) Gex )N [26] alloys. Invar
generally refers to a material with a vanishingly small thermal expansion near room
temperature due to its large positive spontaneous volume magnetostriction [27, 28].
Elinvar refers to a material with a vanishingly small temperature variation of its
elastic properties with temperature. For further discussion of magnetovolume and
invar effects, see [23, 25, 29, 28, 30].

Villari Effect

The Villari effect describes a change of magnetic susceptibility due to straining the
sample [31] by an external stress. It is easier to magnetize a material with positive
magnetostriction in the direction of a tensile stress. The stress-induced lattice strain
induces a change of the magnetic configuration such that those domains grow in
size which have their local magnetization oriented parallel with the lattice strain.
This change of domain configuration affects the susceptibility and the magnetization
curve below saturation. The impact of stress on the magnetic anisotropy of a sample
is discussed in section “Magnetostriction and Stress-Induced Magnetic Anisotropy”.
The stress-induced change of domain configuration is schematically illustrated in
Fig. 8.

E Effect

A material which is not magnetically saturated is composed of magnetic domains.


The magnetic domains are characterized by a saturated magnetization with a
constant magnetization direction within each domain. This magnetization direction
changes from domain to domain. Different domains in a cubic ferromagnet with
[100] anisotropy are separated by 90◦ and 180◦ domain walls. This description is
a valid approximation in many cases, but it should not disguise the fact that more
complex situations such as vortex structures sometimes exist [9, 33].
The application of a tensile stress to a material with positive magnetostriction
rotates the magnetization parallel to the stress direction, and corresponding domains
grow in size. The magnetostrictive strain λ then adds to the elastic strain, and the
material appears softer. The application of a tensile stress to a material with negative
magnetostriction aligns the magnetization perpendicular to the stress direction. A
554 D. Sander

positive magnetostrictive strain λ⊥ still appears along the stress direction, and the
material appears again softer. Thus, irrespective of the sign of magnetostriction, a
magnetic material with domains that are non-collinear with the applied stress in zero
field always appears elastically softer during a magnetization process as compared
to a sample which is magnetically saturated. This variation of elasticity is termed
E effect.
Figure 2 presents an example of the E effect for Ni [32]. The Young modulus
E is affected by the magnetization state of Ni below the Curie temperature of TC =
360 ◦ C. The demagnetized sample shows a considerably reduced Young modulus of
178 GPa below TC , as compared to the magnetically saturated sample with 211 GPa.
Here, a relative reduction of the Young modulus of (Esat − Emagn=0 /Emagn=0 =
0.185) is observed.
The magnitude of the E effect depends on numerous factors. The crystal-
lographic orientation of the sample, its magnetic domain structure, magnitude
and orientation of the applied strain (stress), and the direction and magnitude
of an applied magnetic field need to be considered for a quantitative description
of the effect [32, 4, 10]. The resulting complexity is alleviated by considering
approximations such as a magnetic domain orientation orthogonal to the stress
direction, isotropic elastic and magnetic properties, and minute strains in the
realm of acoustic vibrations. Under these assumptions, the reduction of the Young
modulus due to magnetostrictive contributions to strain is approximated by [10]

Eel − Eλ 9λ2S H 2
= Eel . (1)
Eλ μMS HS3

Here, λS , H , HS , and MS are the saturation magnetostriction, the applied magnetic


field, the magnetic anisotropy field, and the saturation magnetization, respectively.
The
√ E effect manifests as a change of the resonance frequency, which scales
as E, and can be tuned by an applied magnetic field. The effect is strong for

Fig. 2 E effect. Young


modulus E of Ni for a 220
magnetized sample
(M = MS , upper curve
(blue)) and for a 210 M=MS TC
E (GPa)

demagnetized sample
(M = 0, lower curve (red))
[32]. The Young modulus of 200
the demagnetized sample is
smaller than that of the fully
190
magnetized sample below the
Curie temperature TC of Ni of
360 ◦ C 180 M=0
100 200 300 400
temperature (oC)
11 Magnetostriction and Magnetoelasticity 555

Fig. 3 Data of the E effect


obtained from the flexural
vibration of a reed [34]. The
0.8
reed has been annealed in
fields transversal and
longitudinal to its length. The 0.6 transversal

(Eel-Eλ)/Eλ
applied field is oriented along
the length of the reed. The
strain induced by the flexural
0.4
vibration leads to a stronger
reduction of elasticity for
transversal domain
orientation 0.2
longitudinal

0
0 1 2
applied field (mT)

amorphous ferromagnetic metals with vanishing magnetic anisotropy and sizeable


magnetostriction. Here, a relative reduction of E by 45% has been reported [34],
as indicated in Fig. 3. Pronounced magnetic anisotropy opposes the effect, as the
stress-induced reconfiguration of magnetic domains is hampered.
Figure 3 presents an example for the relative variation (Eel − Eλ )/Eλ as a
function of an applied field along the length of a reed. The reduction of the Young
modulus is more pronounced for a sample with magnetic domain orientation along
the transversal direction (along the width of the reed) in zero field as compared to a
sample with longitudinal magnetic domain orientation (along the length of the reed)
in zero field. The transversal domains are rotated in the magnetizing field toward
the longitudinal direction, and this makes their further alignment susceptible to the
vibration-induced lattice strain. The effect vanishes with increasing field, where Eλ
approaches Eel .
The E effect affects elasticity only in the regime of minute strains of the order
of magnetostriction. Thus, typically lattice vibrations and small amplitude flexural
vibrations are influenced, but not elasticity in the range of larger strains of order 1%.

Magnetomechanical Damping

There is a loss of energy associated with any elastic vibration of a sample. Exper-
iments show that this loss is larger for ferromagnetic than for non-ferromagnetic
samples [4, 6]. The reason is that vibrations lead to flexural strains which interact
with the magnetic configuration via magnetoelastic coupling. This link to mag-
netization opens a further loss channel, and a stronger damping results. This is
advantageous for applications where a strong intrinsic damping is required to avoid
catastrophic vibrational amplitudes near resonance, such as in turbines, but it is
556 D. Sander

disadvantageous for applications in ultrasonic transducers, where it waists energy


[36]. The magnetomechanical loss is larger for samples with large magnetoelastic
coupling, and it increases with increasing vibrational amplitude.
The impact of the vibration-induced strain on the magnetic domain configuration
is related to the Villari effect. As a result, the local magnetization of an unsaturated
sample changes during vibration, which gives rise to local eddy current and
magnetomechanical losses. The former increases with increasing frequency of
vibration, whereas the latter is independent of frequency, and it is the prevailing
loss channel at low frequencies in the range of Hz.
The flexural vibration-induced change of magnetization is an hysteric process,
which involves irreversible displacements of domain walls. This leads to losses
in the vibrational energy, and the sample vibration is damped. The magnitude of
damping depends in a subtle manner on the magnetization state of the sample.
A sample with zero net magnetization in zero field shows a larger damping as
compared to a magnetically saturated sample.
This situation is illustrated in Fig. 4. The curves show the amplitude of a torsional
vibration of a 1 mm Fe wire in an axial magnetic field (top curve, blue) and without
an axial magnetic field (bottom curve, red) [35]. The torsion-induced magnetization
change is suppressed by the applied field, and a lower damping results as compared
to the field-free case.

Wiedemann Effect

The Wiedemann effect [37] is the torsional twist observed in a sample upon
the application of an helical magnetic field. The helical field results from the
superposition of an axial magnetic field produced by a surrounding coil and a
circular field produced by a current passing through a wire-shaped or thin-walled
tube sample. The situation is schematically illustrated in Fig. 5, and the resulting
sample torsion γ is indicated. The effect is the basis of a most sensitive measurement
of magnetostriction, as the method can be applied in resonance with a torsional

Fig. 4 Decay of the torsion


torsion amplitude (arb. units)

amplitude of a axial field 10 mT


1-mm-diameter Fe wire with
(top, blue curve) and without 0
(bottom, red curve) an
applied magnetic field along
the wire axis [35]. The
magnetic field suppresses no axial field
magnetomechanical damping
by reducing torsion-induced 0
magnetization changes of the
wire
0 4 8 12 16 20 24
time (s)
11 Magnetostriction and Magnetoelasticity 557

Iwire
γ

Icoil

Fig. 5 Schematic of the Wiedemann effect. A coil surrounding the sample gives an axial magnetic
field along the sample axis. The current through the wire (sample) produces a radial field, which in
conjunction with the axial field results in an helical magnetization of the sample. Magnetostriction
in the sample causes a sample twist, which induces a torsion γ , as depicted in the schematic of the
cross section

vibration, whereby the resulting sample twist is largely amplified. A sensitivity of


10−13 for the measurement of magnetostriction has been reported [38].

Matteucci Effect

The converse to the Wiedemann effect is the Matteucci effect, the appearance of a
helical magnetization component at the end of a twisted sample upon changing the
axial sample magnetization [39]. The effect has been exploited in micro-fluxgate
sensors, where a helical magnetization is imposed on an amorphous Fe77.5 Ni7.5 B15
wire of 0.125 mm diameter. A field sensitivity of order 100 nT has been reported
[40].

Magnetic Field-Induced Strain Phenomena, Which Differ from Joule


Magnetostriction

Joule magnetostriction results from the strain dependence of the magnetic


anisotropy. It is observed as a field-driven change of the linear sample dimensions.
However, there are other phenomena which describe a change of linear dimensions
of a sample for an applied magnetic field [41]. Also there, the term magnetostriction
is used, irrespective of the different physical origin. Here we mention three such
phenomena. We treat shape-memory alloys, where a magnetic field drives a large
strain in the martensitic phase of a material with two twin variants. Huge strains of
order several percent are observed. We also present results on field-driven spin-state
transition in Co3+ and on the interactions between a magnetic field and pinned
Abrikosov flux lattices in type-II superconductors. These effects give strains of the
order of 2 × 10−3 and 2 × 10−4 , respectively.

Ferromagnetic Shape-Memory Alloys (FSMA)


The ferromagnetic shape-memory effect is observed for materials, which undergo
a diffusionless martensitic phase transition between a high-temperature cubic phase
558 D. Sander

and a low-temperature phase of lower symmetry. The dominant lattice distortion


is a shear strain. Sample strains in these phase transitions can be of order several
percent. Such a large strain in the martensitic phase is accommodated by the
formation of twin variants [42]. This means that structural domains with different
c-axis orientations, the twin variants, are characteristic of these materials. The easy
magnetization direction is along the c-axis.
Field-induced strain may be triggered by a field-driven phase transition or by
a field-driven strain developing exclusively in the martensitic phase, where the
materials, in their martensitic phase, show a pronounced dependence of lattice strain
on magnetic field, and field-driven strains of order several percent are possible. The
huge magnetic field-induced strain arises as the Zeeman energy favors the growth
of one variant at the expense of the other. This strain occurs fully within in the
martensitic phase. This field-induced strain is schematically depicted in Fig. 6.
The magnetic field strains in the martensitic phase are due to field-induced twin-
boundary motion [42, 43, 10, 11]. As a result of the formation of crystallographic
twins in the martensitic phase, the magnetization direction varies from one twin
variant to the other. Provided that the magnetic anisotropy of the system is suffi-
ciently strong, an applied magnetic field will not simply rotate the magnetization.
Instead, the growth of the twin variant with favorable magnetization direction with
respect to the applied field is favored, and sizable strain results. The driving force
for the elimination of one twin variant and the growth of another is the minimization
of the Zeeman energy in the applied magnetic field.

cubic
T > Tm
cooling heating
T < Tm T > Tm

H
εH
H=0 applying
a field H

Fig. 6 Illustration of the ferromagnetic shape-memory effect. The parent material is cubic above
the martensitic temperature Tm . A martensitic phase transition occurs upon cooling below Tm . A
lattice strain of order 1% is observed, which is accompanied by the creation of a twin structure
with two variants. The application of a magnetic field H drives the growth of one variant, and a
field-induced strain H is observed. Heating above Tm restores the material to its cubic phase. The
white arrows indicate the magnetization direction within the variants
11 Magnetostriction and Magnetoelasticity 559

The effect is observed in the Heusler alloy Ni2 MnGa. This material has a
martensitic transition between a high-temperature cubic phase and low-temperature
tetragonal phase. The transition temperature is near Tmart = 276 K. The dominant
mechanism of this first-order phase transition is the occurrence of shear strains in
the low-temperature phase. The tetragonal phase is characterized by a huge lattice
strain along the c-axis of order −7%, whereas the lateral strain along a and b is only
of order +1.3% with respect to the parent cubic phase.
It is possible to move the phase transition temperature Tmart to higher values by
modifying the electron-per-atom ratio by alloying [11]. Strains of up to 10% are
achieved for the transition into a orthorhombic phase [44].
It has been found that the temperature dependence of the Ni 3d states plays a
crucial role in the electronic origin of the ferromagnetic shape-memory effect [45].
This supports the view that the martensitic instability can be described as a Jahn-
Teller effect [45].
The ferromagnetic shape-memory effect differs from Joule magnetostriction due
to its different physical origin; no strain dependence of the magnetic anisotropy
is necessary. Also, the effect shows a threshold field, below which no strain is
observed. In addition, the magnetic field-induced strain shows a large hysteresis,
whereas magnetostriction in hard axis magnetization is an anhysteretic process. The
field-induced strain in FSMAs is usually observed only in a narrow temperature
range below Tmart and above a lower cutoff temperature [11].

Magnetic Field-Driven Spin-State Transition in


La(1−x) Srx CoO3−δ (x ≥ 0.3)
Large magnetic field-induced strain of order 1 × 10−3 has been observed Co3+
compounds such as LaSrCoO3 [46] and LaCoO3 [47], at low temperature (25 K)
and in high fields (14 T). This field-driven strain increases monotonically with field,
without saturating. It reaches 2.2×10−3 at 14 T. It has been ascribed to a field-driven
spin transition between a low-spin state and a Jahn-Teller distorted intermediate-
spin state [46]. This effect does not rely on the ferromagnetic state of the sample,
and this contrasts with Joule magnetostriction.

Magnetostriction in Superconductors
The application of a magnetic field along the c-axis of the high Tc superconductor
Bi2 Sr2 CaCu2 O8 in its superconducting state at 5 K leads to a transverse magne-
tostriction in the a, b plane of −2 × 10−4 at 6 T [48]. The negative sample strain
in the plane perpendicular to the field direction is ascribed to a compressive stress,
exerted by pinned Abrikosov vortices which are pushed together by an increasing
field.
After this overview of effects due to magnetoelastic coupling and magnetic field-
induced strain, we now focus on Joule magnetostriction. Relations useful for the
description and measurement of magnetostriction and magnetoelastic coupling in
cubic and hexagonal bulk samples and films are derived in the following sections.
560 D. Sander

Magnetoelasticity and Joule Magnetostriction

We start from the expressions for the elastic fel and the magnetoelastic energy
density fme for a sample with a cubic lattice. These expressions can be found in
many textbooks [7, 8, 10, 12] and reviews [2, 17]. We use the tensor notation for the
lattice strain ij [49]. Confusion is possible because there are different conventions
for representing strain. The contracted Voigt notation and engineering strains are
alternative descriptions, but they give different expressions for the off-diagonal
strain. The different conventions are discussed in the Appendix.

Derivation of the Magnetostrictive Strain Tensor: Cubic Case

The magnetoelastic strain tensor me gives the three-dimensional strain state of a
sample upon saturation magnetization along the direction (α1 , α2 , α3 ), where the
αi are the direction cosines of the magnetization with respect to the cubic axes.
The magnetization-induced strain always refers to a reference state of zero net
magnetization, as given by a demagnetized sample. This demagnetized reference
state is formally implemented by assuming a random orientation of the magnetiza-
tion. By convention, this reference state is characterized by zero magnetostrictive
strain. This convention gives rise to terms of −1/3 in the expression of the αi2 ,
reflecting the average of cos2 (ϑ) over the unit sphere (see Eq. (17)). Note that such
a demagnetized reference state may be hard to realize experimentally. Irrespective
of this, the resulting expressions can be used also for calculating the change of strain
between two well-defined magnetization directions as outlined below. The role of
the reference state for magnetostriction measurements is illustrated in Fig. 7. The
elastic energy density

1
fel = 2
c11 (11 +22
2
+33
2
)+c12 (11 22 +11 33 +22 33 )+2c44 (23
2
+13
2
+12
2
)...
2
(2)
and the magneto elastic energy density

1 1 1
fme = B1 (11 (α12 − ) + 22 (α22 − ) + 33 (α32 − )
3 3 3
+2B2 (23 α2 α3 + 13 α1 α3 + 12 α1 α3 ) . . . (3)

are given in second and first order in strain , respectively [17]. The coefficients
of the rank two strain tensor are ij . The elastic stiffness constants are given by cij
(units N/m2 ), and the magnetoelastic coupling coefficients are given by Bi (units
J/m3 ). Note that this equation is given in a short-hand notation for the cij which
disguises that the elastic stiffness is described by a rank four tensor. Appropriate
tensor transformations need to be applied for the study of directional dependencies
[17]. The dots indicate that higher-order terms are omitted, but may be important
11 Magnetostriction and Magnetoelasticity 561

(a) εII ε
εII
εT

Mavg=0 M=MS 2 S
0
H=0 H=HS ε T

0 HS H II εII
(b) εII 1 Δε 2

ε
T

Mavg=0 M=MS
H=0 H=HS

Fig. 7 (a) Starting from a demagnetized state with a random magnetization configuration, satura-
tion along the field direction gives the lattice strain  along that direction and ⊥ perpendicular to
it. (b) Note that a demagnetized state with a collinear domain orientation along the field direction
does not show any magnetostrictive strain change upon saturation. To avoid the impact of the
often undefined reference state at zero field, a strain measurement during the reorientation of
the magnetization direction between two saturated states 1 (vertical) and 2 (horizontal) gives the
change of strain as  = 32 λS

[17, 50]. This will be discussed below for fme , where the Bi are found to be strain
dependent.
We also note that an alternative description exists within the symmetry-invariant
notation [10]. Here, the free energy expression is expanded in a set of orthogonal
harmonic functions. For a thorough discussion, see [51, 52, 8].
The components of the magnetoelastic strain tensor are obtained by calculating
the strain derivatives of the total energy density ∂f/∂ij with f = fel + fme ,
setting the expressions to zero, and solving the resulting six equations for ij . Some
simplification is possible by exploiting α12 + α22 + α32 = 1, and the following
magnetoelastic strain tensor results:

⎛   ⎞
B1 α12 − 13
⎜ − c11 −c12 − B22cα44
1 α2
− B22cα44
1 α3

⎜   ⎟
⎜ B1 α22 − 13 ⎟
me = ⎜ − B2 α1 α2 − B2 α2 α3 ⎟ .
− 2c44 ⎟ (4)
⎜ 2c44 c11 −c12  
⎝ B1 α32 − 13

B2 α1 α3
− 2c44 − B22cα44
2 α3
− c11 −c12
562 D. Sander

The magnetostrictive strain δ , which is observed in the direction β = (β1 , β2 , β3 ),


where the βi are direction cosines of the strain measurement direction with respect
to the cubic axes, is calculated from

3 3
= βi βj meij (5)

i=1 j =1

B1 1 1 1
=− α12 − β12 + α22 − β22 + α32 − β32
c11 − c12 3 3 3
B2
− (α1 α2 β1 β2 + α1 α3 β1 β3 + α2 α3 β2 β3 ) . (6)
c44

This equation gives zero strain for a state with random magnetization direction.
We calculate the magnetostrictive strains δ for different magnetization and strain
measurement directions. This defines recipes for extracting the tabulated magne-
tostrictive coefficients λ100 and λ111 . We refrain from calling the λ constants, as
their value may be changed due to lattice strain, or in the monolayer thickness range
[53,54,55,56,17,57,58,50]. The first remark refers to a potential strain dependence
of the magnetoelastic coupling coefficients Bi and the second to the contributions
of interface effects to Bi in atomically thin samples.
Magnetization Along [100]; Strain Measurement Along [100]
This defines a procedure to measure λ100 , starting from the demagnetized state with
random domain orientation. With α = (1, 0, 0) and β = (1, 0, 0), we obtain

δ 2B1
= λ100 = − . (7)
3(c11 − c12 )

Magnetization Along [111]; Strain Measurement Along [111]


This defines a procedure to measure λ111 , starting
√ from√the demagnetized
√ state with
random domain orientation. With α = β = (1/ 3, 1/ 3, 1/ 3), we obtain

δ B2
= λ111 = − . (8)
3c44

We see that a negative magnetoelastic coupling coefficient Bi induces a positive


lattice strain. This is plausible by realizing that the energy density f of the system
is lowered by the magnetostrictive strain. Thus, a reduction of f calls for opposite
signs for B and , as found above.
11 Magnetostriction and Magnetoelasticity 563

Finally, we obtain the magnetostrictive strain tensor λ by replacing the expres-


sions of Bi in Eq. (4) with the definitions of λ in Eqs. (7) and (8).
⎛   ⎞
2 λ100 α1 − 3
3 2 1 3 3
⎜ 2 λ111

α1 α2
 2
λ111 α1 α3

⎜ ⎟
λ = ⎜ 3
2 λ111 α1 α2
3
λ
2 100 α 2 − 1 3
λ
2 111 α α
2 3 ⎟ (9)
⎝ 2 3  ⎠
2 λ100 α3 − 3
3 3 3 2 1
2 λ111 α1 α3 2 λ111 α2 α3

δ
We can now express the magnetostrictive strain in terms of the magnetostriction
coefficients λ.

3
= βi βj λij (10)

i=1 j =1
  1
3
= λ100 α12 β12 + α22 β22 + α32 β32 −
2 3
+3λ111 (α1 β1 α2 β2 + α1 β1 α3 β3 + α2 β2 α3 β3 ) (11)

We apply this expression first to derive the difference between longitudinal and
transversal magnetostrictive strain, namely, strain along the magnetization direction
100 −  100
and perpendicular to it. To this end, we calculate the difference of strain 100 010
for a fixed magnetization direction along [100] and strain measurements along [100]
and along the perpendicular direction [010].
Magnetization Along [100]; Strain Measurement Along [100] and [010]
We introduce superscripts and subscripts to identify the magnetization direction and
the strain measurement direction, respectively.

1 3
100
100 = λ100 ; 010
100
= − λ100 ; 100
100
− 010
100
= λ100 (12)
2 2

It follows that λ100 can be derived from two strain measurements along and
perpendicular to the saturation magnetization direction, irrespective of a reference
state with unknown magnetization distribution. We also see that the transversal
magnetostrictive strain is opposite in sign and half in magnitude of the longitudinal
magnetostrictive strain. The same result is obtained for the strain along [001]. Thus,
the sample volume, given by the trace of the strain tensor, remains constant.
The values for λ111 are accessible from measurement along other directions.
Magnetization Along [110]; Strain Measurement Along [110] and [1–10]

1 3 1 3 3
110
110 = λ100 + λ111 ; 1−10
110
= λ100 − λ111 ; 110
110
− 1−10
110
= λ111 (13)
4 4 4 4 2
564 D. Sander

Magnetostriction of Polycrystalline Cubic Materials

We may regard an isotropic polycrystalline sample as a polycrystal with random


orientation of crystallites. In this case, the distinction between λ100 and λ111 is
obsolete, and we introduce λs instead. Such a sample should also have vanishing
elastic anisotropy, and this is reflected by c44 = (c11 −c12 )/2; see Eqs. (7), (8). Thus,
also a single magnetoelastic coupling coefficient B is sufficient for the description.
A compact expression of the magnetostrictive strain of the isotropic sample follows
from Eq. (11) as

δ 3 1
= λs cos2 (ϑ) − . (14)
2 3

The angle ϑ is measured between the magnetization direction and the strain
measurement direction. The reference to the crystal axes is lost.
Determination of the Isotropic Magnetostriction
A strain measurement along the direction of magnetization, λ , and perpendicular
to it, λ⊥ , gives

3
λ − λ⊥ = λs , (15)
2

as is depicted in Fig. 7a.


We can relate the isotropic magnetostriction λs to λ100 and λ111 , if the directional
dependencies of the latter two are considered. To this end, we revert to Eq. (11) with
βi = αi and obtain
 
δ 3 1
= λ100 α1 + α2 + α3 −
4 4 4
+ 3λ111 α12 α22 + α12 α32 + α22 α32 . (16)
2 3

The directional dependence is given by terms cos2 (ϑ), cos4 (ϑ), and cos2 (ϑ)
sin2 (ϑ). A random orientation of crystallites, which is the basis of the isotropic
sample described here, is considered by averaging the directional dependence over
the unit sphere. We calculate the respective averages

2π π 1
cos (ϑ)avg =
2
cos2 (ϑ) sin(ϑ)dϑdϕ/(4π ) =
0 0 3
2π π 1
cos4 (ϑ)avg = cos4 (ϑ) sin(ϑ)dϑdϕ/(4π ) =
0 0 5
2π π 2
cos2 (ϑ) sin2 (ϑ)avg = cos2 (ϑ) sin3 (ϑ)dϑdϕ/(4π ) = (17)
0 0 15
11 Magnetostriction and Magnetoelasticity 565

and finally obtain the isotropic magnetostriction λs

2 3
λs = λ100 + λ111 . (18)
5 5

This expression was first presented by Akulov [59] and was later criticized
for neglecting elastic anisotropy [60], which leads to a correction, which can be
negligible or sizable, depending on the material [60].

Derivation of the Magnetostrictive Strain Tensor: Hexagonal Case

We follow the same procedure as for the cubic case above. We start from the
following expressions for the elastic and the magnetoelastic energy density [17].
We choose a coordinate system where the z-axis coincides with the c-axis and the
x-axis with one of the unit vectors of the basal plane. The reference state is ascribed
to magnetization parallel to the c-axis ( α3 = 1, α1 = α2 = 0).

1 1
fel = 2
c11 (11 + 22
2
) + c33 33
2
+ c12 11 22 + c13 (11 33 + 22 33 )
2 2
+ 2c44 (23
2
+ 13
2
+ (c11 − c12 )12
2
+ ... (19)
fme = B1 (11 α12 + 212 α1 α2 + 22 α22 ) + B2 (1 − α32 )33
+ B3 (1 − α32 )(11 + 22 ) + 2B4 (23 α2 α3 + 13 α1 α3 ) + . . . (20)

We calculate the strain derivative of the total energy density, set it to zero, and
solve for the ij . We then calculate the magnetostrictive strain for four different
magnetization states [61, 62] and obtain the following definitions for the Bi :

B1 = −(c11 − c12 )(λA − λB ); B2 = −c13 (λA + λB ) − c33 λC


B3 = −c12 λA − c11 λB − c13 λC ; B4 = c44 (λA + λC − 4λD ) (21)

For convenience, we present the expressions for λi in terms of the Bj in the


Appendix.
We obtain the following magnetostriction tensor:
⎛ ⎞
λA α12 + λB α22 (λA − λB )α1 α2 λACD α1 α3
λhex = ⎝ (λA − λB )α1 α2 λA α22 + λB α12 λACD α2 α3 ⎠ (22)
λACD α1 α3 λACD α2 α3 λC (1 − α32 )

with the abbreviationλACD = − 12 (λA + λC − 4λD ). This gives the magnetostrictive


strain δ
in the hexagonal system
566 D. Sander

3 3
= hex
βi βj λij (23)

i=1 j =1

= λA (α1 β1 + α2 β2 )(α1 β1 + α2 β2 − α3 β3 ) + λB (α2 β1 − α1 β2 )2


+λC β3 (α1 α3 β1 + α2 α3 β2 + (α32 − 1)β3 )
+λD 4α3 β3 (α1 β1 + α2 β2 ). (24)

The individual magnetostriction coefficients λi are defined by the following


orientations of magnetization α = (α1 , α2 , α3 ) and strain measurement β =
(β1 , β2 , β3 ):

λA : α = (1, 0, 0), β = (1, 0, 0)


λB : α = (1, 0, 0), β = (0, 1, 0)
(25)
λC : α = (1, 0, 0), β = (0, 0, 1)
λD : α = ( , 0, ), β = ( √1 , 0, √1 )
√1 √1
2 2 2 2

Magnetostriction of Polycrystalline Hexagonal Materials

We derive the expression for the effective magnetostriction in a polycrystal with


random orientation of grains with hexagonal symmetry. We distinguish between λ
and λ⊥ for strain measurements along the magnetization direction and perpendicular
to it, respectively. Starting point is Eq. 24, where we set βi = αi to derive λ .

λ = λA ((1 − α32 )2 − α32 (1 − α32 )) + 4λD α32 (1 − α32 )


= λA (sin4 (ϑ) − cos2 (ϑ) sin2 (ϑ)) + 4λD cos2 (ϑ) sin2 (ϑ) (26)

Here, ϑ is the angle between the magnetization direction and the c-axis. With
the calculated averages of the trigonometric functions from Eq. (17) and with
(sin4 (ϑ))avg = 15
8
, we get

2 8
λ = λA + λD . (27)
5 15

The relative volume change of magnetostriction ω is given by the trace of the


magnetostrictive strain tensor of the hexagonal system in Eq. (22). It is linked to
both λ and λ⊥ via

ω = (λA + λB + λC ) sin2 (ϑ) (28)


= λ + 2λ⊥ . (29)
11 Magnetostriction and Magnetoelasticity 567

The volume magnetostriction of a sample with hexagonal symmetry depends on


the orientation of the magnetization with respect to the c-axis. Replacing sin2 (ϑ)
with its average 23 , inserting λ , and solving for λ⊥ give

1
λ⊥ = (2λA + 5λB + 5λC − 4λD ). (30)
15

Thus, the magnetostrictive strain λ , λ⊥ of a sample with random orientation of


crystallites with hexagonal symmetry depends on all magnetostrictive coefficients
λ A . . . λD .

Magnetostriction and Stress-Induced Magnetic Anisotropy

Inspection of Eqs. (3) and (20) reveals that magnetoelasticity is intimately linked to
magnetic anisotropy insofar as it involves terms which depend on the direction of
magnetization. Consequently, the application of stress τ induces a lattice strain ,
and this strain contributes via the magnetoelastic coupling coefficients Bi to the
energy density. It is important to note that it is the stress-induced strain which
drives the corresponding change of magnetic anisotropy. This impact of an imposed
lattice strain on magnetic anisotropy is often referred to as inverse magnetostrictive
effect. Theory exploits the link between lattice strain and magnetoelastic effects
by identifying magnetoelasticity as the derivative of magnetic anisotropy with
respect to strain [63, 64]. This aspect is elucidated in section “Magnetostriction and
Magnetoelasticity: Physical Origin and Insights from Theory”. Here, we briefly note
that for 3d metals the corresponding physics is due to the spin-orbit interaction. Its
dependence on lattice strain is a daunting topic. This is also due to the small energy
scales, which typically are in the μeV per atom range. Reliable calculations with
the required numerical accuracy remain challenging.
Thus, phenomenological pictures are still used to describe stress-induced mag-
netic anisotropy. Here, we focus on illustrative examples to provide a basic
description of the matter for a cubic system. We calculate the stress-induced lattice
strain and insert the obtained strain in the equation of the magnetoelastic energy
density fel of Eq. (3). Then, the orientational dependence of the expression is
analyzed.
The stress-strain relation is derived from the expression of the elastic energy
density of Eq. 2 by exploiting τij = ∂fme /∂ij . In the following, we neglect shear
stress by setting τij = 0 for i = j and focus on the cubic system. We obtain

τ11 = c11 11 + c12 22 + c12 33


τ22 = c12 11 + c11 22 + c12 33
τ33 = c12 11 + c12 22 + c11 33 (31)

and solve these equations for ii to derive the stress-induced lattice strain.
568 D. Sander

Isotropic Stress
An isotropic stress τ = τii gives rise to an isotropic strain of magnitude  = ii =
τ/(c11 + 2c12 ). An isotropic contribution to fel of magnitude B1 τ/(c11 + 2c12 )
results. Thus, the magnetic anisotropy, defined as energy difference between two
magnetization directions, remains unaffected.
Uniaxial Stress
We assume an uniaxial stress τ11 = τ, τ22 = τ33 = 0. A three-dimensional strain
state results, where the lattice strain differs along the direction of applied stress from
its value in the plane perpendicular to the stress direction. We obtain

c11 + c12
11 = τ
(c11 − c12 )(c11 + 2c12 )
c12
22 = 33 = −τ . (32)
(c11 − c12 )(c11 + 2c12 )

The ratio −22 /11 = −33 /11 defines the Poisson ratio ν = c12 /(c11 + c12 ).
Insertion into Eq. 3 and omitting isotropic terms give the stress contribution to
magnetic anisotropy as

B1
τ
fme =τ α2
c11 − c12 1
3
= − τ λ100 cos2 (ϑ), (33)
2

where ϑ gives the angle between the direction of stress and magnetization. Thus, an
uniaxial magnetic anisotropy results.
How large are stress-induced magnetic anisotropies? Can they potentially change
the easy magnetization direction as given by the magnetocrystalline anisotropy?
Stress-induced magnetic anisotropies scale as Bi . Typical values for B are of order
several MJ/m3 . The strain in bulk samples is certainly always below 0.01, and thus
the stress contribution in a bulk sample can reach 0.1 MJ/m3 . This is a sizable
contribution to magnetic anisotropy in comparison to K1 = 0.05 MJ/m3 of bulk
Fe [17, 65, 58]. For systems with smaller magnetocrystalline anisotropy, such as
Ni or amorphous metals, stress-induced magnetic anisotropy is often the dominant
source. In epitaxial ultrathin films, much larger strains of up to several percent may
prevail, and stress-induced anisotropy should always be considered [17, 65].
In view of the impact of magnetoelasticity on anisotropy, it is important to realize
that terms B, Bτ, τ λ contribute. Thus, the application of a tensile stress τ > 0 to
a sample with positive magnetostriction (λ > 0, B < 0) has the same impact on
anisotropy as the application of a compressive stress (τ < 0) to a sample with
negative magnetostriction. If λ is positive, then it is easier to magnetize the material
in the direction of a tensile stress. This link between stress and magnetic anisotropy
plays a role for the domain distribution in stressed materials and for the approach to
11 Magnetostriction and Magnetoelasticity 569

(a) λ>0

τII τII
MII
τII, λ>0
MS
V(MII)>V(M ) τII, λ<0
τ=0
T

(b) λ<0

Mavg=0
τII τII
τ=0 0
V(M )>V(MII)
T
0 H II τ
Fig. 8 The impact of stress on the magnetization curve of a magnetostrictive material. (a) Tensile
stress favors a longitudinal domain configuration in a material with positive magnetostriction. (b)
A transversal domain configuration is favored for negative magnetostriction. Magnetization along
the tensile stress direction gives a reduced effective anisotropy in (a) (red curve) and a larger
anisotropy in (b) (blue curve)

saturation magnetization. The interplay between stress and domain configuration of


a sample is schematically illustrated in Fig. 8. The interplay between stress, strain,
and domain configuration is also relevant for the E effect discussed above.

Magnetoelastic Effects in Films

Magnetoelastic effects are of significant importance for understanding magnetic


anisotropy in thin films [17, 58, 65, 66, 50]. The reason is that magnetoelasticity
couples the lattice strain to the magnetization direction via the coefficients Bi ,
as expressed in Eqs. (2) and (20). Films are, as result of the deposition on a
substrate, almost always strained. The situation is often transparent for epitaxial
pseudomorphic films on single crystal substrates, where the misfit between film and
substrate gives the lattice strain [67, 68, 69]. The resulting film strain is anisotropic.
It differs in sign and magnitude for the in-plane and the out-of-plane direction [17].
This three-dimensional strain state of the film modifies the magnetic anisotropy.
Easy magnetization directions, which differ from those of the respective bulk
materials, are commonly observed. Magnetoelasticity is often the key driving force
[17, 65].
Numerous experiments have established that magnetoelasticity in films may
deviate from the bulk [53, 54, 57, 55, 56, 41, 17, 58, 50]. The sign and magnitude
of Bi are found to change with lattice strain and film thickness. A theoretical
understanding of this phenomenon is progressing [70, 71, 72, 73, 74, 75, 76, 77].
Theory provides physical insights in support of a strain-dependent correction

B1eff = B1 + D1 ij , (34)


570 D. Sander

leading to an effective magnetoelastic coupling coefficient B1eff . The magnitude


of the correction term D1 is sizable, of order 100–1000 MJ/m3 . Table 4 provides
an overview of corresponding experimental and theoretical data. Thus, strains
in the percent range may already change the effective magnetoelastic coupling
considerably.

Experimental Determination of Magnetostriction and


Magnetoelastic Coupling

A direct measurement of magnetostriction in bulk samples is possible by measuring


the change of length of a sample upon magnetization as described by Eqs. (11) and
(24). The resulting sample distortion is schematically indicated in Fig. 9.
Strain measurements by strain gauges, by capacitive and optical measurements,
have been described to measure the usually small relative change of length of order
10−6 with high precision [78]. Also, the variation of the tunnel current between
tip and magnetostrictive sample, as induced by a magnetostrictive change of length
of the tunnel barrier, has been analyzed to derive magnetostriction data [79, 80].
It appears that among all techniques the capacitance method, where the change
of length of the sample induces a change of distance of the plate separation in
a capacitor, has the highest sensitivity and reliability. It is a bulk, rather than a
surface measurement where the sample is part of the electrode of a parallel-plate
capacitor, and its change of capacity is measured with a three-terminal capacitance
measurement [81, 82, 83, 84, 85, 86, 87, 88, 89]. Thus, even minute magnetostriction
of order 10−10 in paramagnetic samples has been detected [90].
The accurate measurement of magnetostrictive strain requires the consideration
of the form effect. The magnetization of a sample leads to a demagnetizing field
within the sample volume which is related to the demagnetizing factor of the

(a) (b)

Fig. 9 (a) Magnetostriction in bulk samples is accessible from the magnetization-induced change
of the lattice strain λ , λ⊥ . (b) Magnetization of a film induces a biaxial magnetoelastic stress,
which is related to the magnetoelastic coupling coefficients Bi . A thin substrate undergoes a
anticlastic curvature with radii of curvature of opposite sign along and perpendicular to the
magnetization direction. Curvature measurements give Bi ; see Eq. (36)
11 Magnetostriction and Magnetoelasticity 571

sample shape. Thus, for magnetostatic reasons alone, a magnetized sample tends
to elongate along the magnetization direction, thereby lowering its demagnetizing
energy. This leads to the superposition of the form effect-driven lattice strain with
the magnetostrictive lattice strain. When λ is small, both can be of comparable
magnitude, and corresponding corrections are required [91].
The Wiedemann Effect: Magnetization-Induced Sample Twist
Alternatively, the Wiedemann effect can be applied to deduce magnetostriction from
a magnetization-induced sample twist (Fig. 5). The helical twist is straightforwardly
measured with high sensitivity by reflecting a laser beam from the sample onto a
position-sensitive detector. This method is extremely useful for samples in the form
of thin ribbons and wires [92,38], where strain measurements are otherwise difficult
to implement. Quantitative analysis reveals [92] a non-monotonic dependence of the
torsion per unit sample length ξ on the magnetic field produced by coil and current.
The maximum torsion per length is related to the isotropic magnetostriction constant
λs via

2.2λs
ξmax = , (35)
a

where the sample thickness is 2a.


Indirect Methods to Determine Magnetoelastic Properties
Magnetoelasticity also reveals itself by its impact on the magnetization process.
Thus, indirect access to magnetoelasticity is possible from an analysis of magne-
tization curves [53, 54]. Straining a magnetostrictive sample modifies its magnetic
properties. Magnetoelasticity couples applied stress with magnetic anisotropy and
magnetic susceptibility. The contribution of an externally applied stress to the
magnetic anisotropy of the system is given by Eq. (33). The resulting change of
the effective magnetic anisotropy is accessible from the change of magnetization
curves and from the modified initial susceptibility [54]. Straining an amorphous
sample with negative magnetostriction increases the effective magnetic anisotropy
and lowers the initial susceptibility.
Other techniques which have the potential to detect changes of the magnetic
anisotropy upon the application of stress, such as small angle magnetization rotation
(SAMR) [93, 94], ferromagnetic resonance (FMR) [95, 96], and Brillouin light
scattering (BLS) [97], give access to magnetoelastic coupling. The analysis of the
data requires modeling of the contributions to magnetic anisotropy and its strain
dependence to extract the magnetoelastic coupling coefficients reliably.

Magnetoelastic Coupling in Films

The indirect methods described above may also be used to study magnetoelastic
effects in thin films. However, they lack the direct link between the experimental
572 D. Sander

observable and the magnetoelastic coupling coefficients that exists in magnetoelas-


tic stress measurements, which are described next.
Magnetostriction is a property of bulk samples which are free to strain upon
changing the magnetization state. Ferromagnetic films are usually bonded to a
substrate, which inhibits film strain upon changing the film magnetization. Instead,
a magnetoelastic stress arises, and the resulting strain depends on the stiffness of
the film-substrate composite. This magnetoelastic stress is of magnitude Bi , and
it amounts to several MPa. Although it is three orders of magnitude less than
typical epitaxial misfit stress, which amounts to GPa for a misfit in the 1% range
[17]. The magnetoelastic stress can still be detected by the stress-induced curvature
of the film-substrate composite. Figure 10 gives a schematic of a corresponding
experimental arrangement which is used to prepare ferromagnetic monolayers and
analyze their stress and magnetic properties under ultrahigh vacuum conditions in
situ [50, 98].
Figure 10 provides a sketch of an optical two-beam curvature measurement
in conjunction with a magnet system for magnetoelastic studies under ultrahigh
vacuum conditions. The optical deflection measurement is mounted to a window

Fig. 10 (a) Sketch of an ultrahigh vacuum chamber for in situ film growth with magnets for
magnetoelasticity studies. (b) Optical curvature measurement. The optics is mounted outside of the
vacuum chamber, and the sample is in the vacuum chamber. 1: sample manipulator, 2: substrate
with film (3) on front surface, 4: laser diode with optics, 5: beam splitter, 6: mirrors, 7: piezo
transducer, 8: split photodiode. (c) Schematic curvature change upon an in-plane magnetization
reorientation of a film-substrate composite with positive Beff
11 Magnetostriction and Magnetoelasticity 573

flange of the vacuum chamber for in situ stress measurements during film growth
and during magnetization processes. The experiment involves a measurement of the
change of substrate curvature upon an in-plane reorientation of the magnetization
direction [17, 65, 99]. The curvature along the sample length is measured for mag-
netization along its length and magnetization along its width. The magnetoelastic
stress produces an anticlastic curvature of the sample.
Let us assume a cubic (100)-film with negative B1 (λ100 > 0). The sample
length is along [100]. Magnetization along [100] induces a compressive stress along
[100], as the film has the tendency to expand along its magnetization direction.
This induces a negative curvature along the sample length. A tensile stress results
along the perpendicular in-plane direction. Magnetization along the film width [010]
rotates the compressive stress to the film width, and a tensile stress along the
film length results. An anticlastic curvature results in both magnetization states.
The sample changes its curvature from curved downward to curved upward upon
magnetization reorientation. This situation is schematically shown in Fig. 10c.
This curvature change is directly linked to the effective magnetoelastic coupling
coefficient Beff [17]
 
Es ts2 1 length 1 width
Beff = − . (36)
6(1 + νs )tf Rx Rx

Note that the elastic properties of the substrate enter via the Young modulus
Es and the Poisson ratio νs . Substrate and film thickness are given by ts and tf ,
respectively. The superscripts give the magnetization direction.
Different magnetoelastic coupling coefficients Bi can be measured depending on
the crystallographic orientation of the film. Table 1 gives an overview of the required
experiments to measure Bi [99].
The quantitative analysis of magnetoelastic curvature measurements needs to
take crystalline anisotropy and substrate clamping into account [100, 99]. The
use of substrates with a length-to-width ration larger than three minimizes the

Table 1 Effective magnetoelastic coupling coefficient Beff for films with the given symmetry and
orientation for an in-plane reorientation of the magnetization. The Bi are defined in Eq. (20). x1
and x2 denote the direction along film length and film width, respectively
Film structure Beff
Cubic, film axes along x1 , x2 B1
Cubic, film axes rotated by 45◦ with respect to x1 B2
Hexagonal, c-axis in-plane,
2 domains parallel to x1 , x2 1
2 (B1 − B2 + B3 )
Hexagonal, c-axis in-plane,
2 domains rotated by 45◦ with respect to x1 B4
Hexagonal, c-axis in-plane,
3 domains rotated by 120◦ with respect to x1 1
4 (B1 − B2 + B3 ) + 12 B4
574 D. Sander

detrimental effect of substrate clamping on the two-dimensional curvature state.


The radius of curvature is in the range 100 m–100 km for magnetoelastic stress
measurements of nm thin films. Magnetization-induced curvature measurements
have been successfully applied to measure magnetoelastic coupling coefficients in
films as thin as a few atomic layers [101, 102, 103, 104, 56, 17].

Magnetostriction and Magnetoelasticity: Physical Origin and


Insights from Theory

Magnetostriction and magnetoelasticity are due to the coupling between magnetism


and atomic structure [63,105,106]. This is brought about by the spin-orbit coupling.
It can be understood as the coupling between the electron spin moment with the
magnetic field created by its own orbital motion. This orbital motion is coupled to
the lattice by the Coulomb potential of the atom cores. As a result, the energy of
the system changes depending on the orientation of the magnetization with respect
to the lattice. This defines the magnetocrystalline anisotropy, and it leads to the
occurrence of specific easy magnetization directions [107, 108, 109, 110, 111, 112,
105, 113, 106, 114].
The magnetocrystalline anisotropy itself depends on lattice strain [106]. Thus,
the total energy of the system with a specific magnetization orientation can
be lowered by a concurrent lattice strain [64]. The strain dependence of the
magnetocrystalline anisotropy is the basis for current calculations of magnetoelastic
effects [63, 115, 76, 113, 106, 116].
Despite the considerable advance of electronic structure calculations, the theoret-
ical study of magnetic anisotropy and magnetoelasticity in the framework of density
functional theory remains a challenging task [73, 106].
Highly accurate numerical calculation of the total energy density ftot of a sample
upon a small lattice strain identifies the magnitude of magnetostriction as the strain
which minimizes ftot . Corresponding calculations for Ni presented in Fig. 11 [64]
reveal a parabolic dependence of ftot on lattice strain. Note the small magnitude
of the energy scale of 0.1 μeV per atom. An amazingly high numerical accuracy is
required to tackle magnetoelastic effects reliably. The calculated magnetostriction
values, as given by the minima of the plot, have the correct sign, but are roughly a
factor four larger than the experimental values (λ100 = −64.5 × 10−6 and λ111 =
exp. exp.

−28.3 × 10−6 ) [17].


The small magnitude of magnetic anisotropy of 3d transition metals of order
several μeV per atom requires high numerical accuracy for the theoretical treatment
of this topic. The magnetocrystalline anisotropy EMCA has been calculated for
bulk metals, 3d films, and nanostructures [117, 113, 118]. Theory also derived the
variation of EMCA with lattice strain, and this property is related to magnetoelas-
ticity. The slope of the plot EMCA as a function of lattice strain is related to the
magnetoelastic coupling coefficients [119, 105].
11 Magnetostriction and Magnetoelasticity 575

Fig. 11 Calculated total


energy density in dependence
of a tetragonal and a trigonal Ni trigonal
-0.05
strain of Ni [64]. The minima
identify the magnetostriction λ111
constants λ100 : −245 × 10−6
and λ111 : −110 × 10−6 .

energy (μeV/atom)
Note the small energy scale
of order 100 neV or 1 mK per -0.15
atom

-0.25 tetragonal
λ100

-200 -100
distortion (10-6)

Fig. 12 Calculated change


of density of states upon
uniaxial compression of bulk 0.04 d z2 d 2 2
Fe from a tight-binding 0.02
calculation. The plot indicates
a shift to lower (higher) 0
energy of bands with dz2
(red) (dx 2 −y 2 ) (blue) -0.02
symmetry. This changes the
spin orbit coupling -0.04
interaction, and consequently
the magnetic anisotropy of -2 1
-3 -1 0
the system is modified [115]

Ab initio calculations offer some insights into the physical origin at the electronic
level of the strain dependence of the magnetocrystalline anisotropy [115]. Figure 12
shows the calculated change of the density of states upon uniaxial compression
of bulk Fe. This tetragonal strain shifts and broadens electronic bands. As a
consequence, the spin-orbit coupling between bands of different symmetry near
the Fermi energy is changed, and this drives a strain-induced change of magnetic
anisotropy [115].
Recent measurements of magnetoelastic coupling in atomic layers have revealed
that film strain may modify the magnetoelastic coupling coefficients in a way that
can be described by higher-order strain contributions to the magnetoelastic energy
density [50]. Calculations [71, 70, 76, 74, 75, 77] support the phenomenological
introduction of strain-modified magnetoelasticity, as discussed in section “Magne-
toelastic Effects in Films”.
576 D. Sander

A comparison between different calculational schemes within density functional


theory and experimental data is presented in Table 4. Inspection of the data reveals
significant quantitative differences between experiment and theory which illustrate
the challenge presented by accurate calculation of magnetostrictive properties.
Nonetheless, recent theoretical work has succeeded in shining fresh light on the
electronic origin of the intriguingly large magnetostriction of some Fe alloys
[120,121,122,116] (see section “Magnetostriction Data of Fe-Ga (Galfenol), Fe-Ge,
FeAl, Fe-Si, Fe-Ga-Al, and Fe-Ga-Ge Alloys”) and of the large positive volume
magnetostriction of invar alloys [28, 30].
The electrons responsible for magnetism in transition metals are largely delo-
calized, and one refers to this class of materials as itinerant magnets. However,
ferromagnetism also exists in 4f metals, albeit with significantly lower ordering
temperatures, and the electronic picture of magnetism is quite different. The origin
of rare-earth magnetism is ascribed to localized 4f electrons. The 4f ferromagnetic
elements show a pronounced magnetic anisotropy, which is ascribed to a strong
intra-atomic spin-orbit coupling of the 4f electrons [123]. This intra-atomic
coupling may exceed the anisotropic interaction with all other charges of the metal
[124]. In the limiting case of infinite anisotropy, one may envision magnetostriction
in 4f metals as being due to the magnetic field-induced rotation of the highly
anisotropic charge distribution of the 4f electrons. Prototypical examples are Tb
and Er. The charge distribution of the former is oblate and that of the latter prolate.
One may expect that the magnetostrictive strain into the direction of a strong
magnetic field differs in sign for both elements. This is indeed calculated and
observed experimentally [123]. Tb expands along the c-axis for magnetization along
the c-axis, whereas Er contracts for the same measurement. In view of the paramount
importance of the 4f charge distribution for magnetism, one refers to single-ion
anisotropy in these elements to distinguish it from the itinerant magnetism of 3d
metals.
Antiferromagnetism and ferrimagnetism are found in oxides of 3d metals.
Superexchange between 3d cations, mediated by O2− anions, drives the magnetic
order. Examples are the antiferromagnetic monoxides MnO, FeO, CoO, NiO, and
ferrites of the general composition MFe2 O4 (M = Cr, Mn, Fe, Co, Ni, Cu, Zn) [10],
respectively. The antiferromagnetic oxides show a lattice distortion below the Néel
temperature TN , which can be described as a magnetostrictive phenomenon in a
wider sense [125]. The discussion of magnetostriction in these oxides and ferrites
relies on the single-ion anisotropy model [125, 126, 10]. A striking example is
CoO, which is cubic (a = 4.261 Å) above TN = 288 K. It contracts at lower
temperature along the spin axis (c-axis) and expands in the perpendicular plane
(ab-plane) [127]. Extrapolation to 0 K gives a strain of magnetic origin along c
by −1.5% and an expansion of +0.9% along a. This large lattice strain has been
ascribed to a repulsive Coulomb interaction between the 3d electrons of Co2+
cations, which are responsible for the magnetism and the charge distribution of
the O2− anions [126]. The regular orthogonal environment seen by the Co-cation,
as produced by the nearest neighbor O-anions, is decisive for an understanding
on the electronic level. A crystal field results, which splits the electronic levels
11 Magnetostriction and Magnetoelasticity 577

of Co2+ into double-degenerate eg states at higher-energy and lower-lying triple-


degenerate t2g states. The orbital moment of Co2+ in CoO is not quenched, and
it is free to align along the spin axis. The corresponding charge distribution is of
pancake shape, with its wider extension in the ab-plane, giving rise to the repulsive
Coulomb interaction in that plane. The Co2+ cation in CoFe2 O4 ferrite of spinel
structure experiences a crystal field of different symmetry. The 3d electronic states
of the cations interact also with next-nearest neighbor Co-cations, in addition to
interactions with nearest neighbor O-anions in the orthogonal arrangement. The
former gives rise to an interaction potential of trigonal symmetry with respect to the
< 111 >-axis. This splits the triple-degenerate t2g state into a lower-lying singlet
and a higher doublet state. The filling of this electronic-level scheme leads also to
an unquenched orbital moment, and a remarkably large magnetostriction at room
temperature of λ100 = −590 × 10−6 results [128, 126]. It is ascribed to the strain
dependence of the crystal field energy.
If the theoretical description of magnetostriction is demanding for bulk samples,
this is even more true for the theory of magnetoelasticity of films [50]. In view
of the electronic origin of magnetic anisotropy and its strain dependence, it is
to be expected that electronic hybridization between film and substrate, or film
and adsorbate layer, modifies not only the atomic structure near the interface
[129, 130] but also spin polarization [131, 132, 129], magnetic anisotropy [65], and
magnetoelasticity [17, 58, 50]. Thus, with decreasing film thickness down to single
atomic layers, the discussion of magnetoelasticity in terms of lattice strain alone
misses important aspects.

Compilation of Data

Magnetoelastic and Elasticity Data for Bulk Transition Metals

Tables 2 and 3 present magnetoelastic coupling and magnetostrictive coefficients,


and elasticity data, respectively, for selected 3d bulk metals.
A recent study on the magnetostriction of binary and ternary Fe alloys [91]
gives values for pure Fe at room temperature and at 77 K (in parentheses) λ100 =
23.3(22.6) × 10−6 and B1 = −3.4(−3.5) MJ/m3 . These values are in close
agreement with the values presented in Table 2.

Table 2 Room temperature values of λ and B [17]. Note that Eq. (3.4) in [17] should read B4 =
c44 (λA + λC − 4λD ). The Bi are calculated from the λi . The required elasticity data are in Table 3.
The values for fcc Co are extrapolated from data for PdCo alloys [17]
Element B1 B2 λ100 λ111 B3 B4 λA λB λC λD
zbcc Fe −3.43 7.83 24.1 −22.7
fcc Co −9.2 7.7 75 −20
fcc Ni 9.38 10 −64.5 −28.3
hcp Co −8.1 −29 28.2 37.4 −50 −107 126 −105
578 D. Sander

Table 3 Elastic stiffness constants cij , Young modulus E in GPa. The latter is calculated for the
(0001) and (100) orientation for hcp and cubic elements, respectively. The Poisson ratio is given
by ν. (Data from [17])
Element c11 c12 c44 c13 c33 E ν
hcp Co 307 165 75.5 103 358 211 0.49
fcc Co 242 160 128 114 0.40
bcc Fe 229 134 115 131 0.37
fcc Ni 249 152 118 133 0.38

Theoretical and Experimental Values of Magnetoelastic Coupling


Coefficients and Their Strain Dependence (Table 4)

Table 4 Compilation of theoretical values and experimental results of magnetoelastic coupling


coefficients Bi and the nonlinear strain-dependent coefficient Di with B eff = Bi + Di  [58].
B1 was derived from the published λ100 with Eq. 7 and the elastic constants from Table 3.
Experimental data for fcc Co are extrapolated from data for PdCo alloys [17]. All calculations
are for bulk under specific strain states. (All values are given in MJ/m3 .) LSDA local-spin-density
approximation, GGA generalized-gradient approximation
B1 D1 B2 2D2 B3 B4 D4
fcc Co
Exp. bulk [17] −9.2 – 7.7 –
LSDA [133] −11.3 – – –
LSDA [73, 74] −15.9 212 3 58
LSDA [113] −13.8 – 10.6 –
GGA [133] −6.9 – – –
GGA [73, 74] −9.8 186 4.5 −71
GGA [113] −5.9 – –
Exp. film Co/Ir(100) [58] 3.5 −842 1.8 930
fcc Ni
Exp. bulk [17] 9.38 – 10 –
LSDA [133] 9.2 – – –
LSDA [73, 74] 12.6 −103 16.9 −132
LSDA [113] 13.9 – 38.8 –
GGA [133] 8.1 – – –
GGA [73, 74] 10.2 −53 11.1 −47
GGA [113] 13.7 – – –
Exp. film Ni/Cu(100) [134] 9.4 −234 10 –
Exp. film Ni/Ir(100) [58] 1.3 273 6.6 −408
bcc Fe
Exp. bulk [17] −3.43 – 7.83 –
LSDA [133] −7.4 – – –
LSDA [73, 74] −10.1 337 −7.0 −40
LSDA [113] −8.3 – −8.9 –
(continued)
11 Magnetostriction and Magnetoelasticity 579

Table 4 (continued)
B1 D1 B2 2D2 B3 B4 D4
GGA [133] −4.1 – – –
GGA [73, 74] −2.4 383 −3.9 18
GGA [113] −4.8 – – –
Exp. film Fe/MgO(100) and −3.4 1100 7.8 −365
Fe/Cr/MgO(100) [57]
Fe/Ir(100) [58] −3.6 155 – –
Fe/W(100) [17] −1.2 200 – –
hcp Co
Exp. bulk [17] −8.1 – −29 – 28.2 37.4
Exp. film Co/W(001) [135] 3.4 1346

Magnetostriction Data of Amorphous Fe Alloys (Table 5)

Table 5 Low temperature Alloy λS , 4.2 K λS , 300 K


and room temperature values
a-Fe80 B20 48 32
of λS in 10−6 for amorphous
alloys [11] a-Fe40 Ni40 B20 20 14
a-Co80 B20 −4 −4

Magnetostriction Data of Fe-Ga (Galfenol), Fe-Ge, FeAl, Fe-Si,


Fe-Ga-Al, and Fe-Ga-Ge Alloys

Magnetostrictive and elastic data of these binary and ternary alloys [136] have
been measured and compiled for different alloy compositions with high resolution
on the composition scale [91]. Table 6 presents specific examples of those alloy
compositions where magnetostriction is much larger or smaller than that of pure Fe.
Note that a tenfold increase of magnetostriction of Fe71.2 Ga28.8 as compared
to Fe is observed, although the magnetoelastic coupling coefficient is only twice
as large as that of Fe. This indicates a contribution of elastic softening to the
remarkably large magnetostrictive strain in this alloy. Equation 7 shows that a
small value of (c11 − c12 ) leads to a large magnetostriction λ100 for a given
magnetoelastic coupling B1 . Ultrasound spectroscopy revealed a decrease of the
tetragonal shear modulus 0.5(c11 − c12 ) from the Fe bulk value of 48 to 6.8 GPa
for a Ga content of 27.2% at room temperature [137]. In addition to lattice
softening, density functional theory has revealed the atomic arrangement of Ga in
Fe-Ga alloys as a further important aspect. The Ga-induced change of electronic
structure is of key importance for the understanding of large magnetostriction
in these alloys [138, 139]. A structural origin of large magnetostriction, due
to nanoscale precipitates of different crystallographic order, has been proposed
[140, 141, 142, 143], but this could not be supported by theory [139].
580 D. Sander

Table 6 Room temperature Alloy λ100 B1


values of λ100 and B1 for
Fe86.6 Ga13.4 143 −12.6
cubic Fe alloys with different
composition [91]. All λ in Fe81.8 Ga18.2 185 −10.6
10−6 , B1 in MJ/m3 Fe71.2 Ga28.8 242 −6.49
Fe89.9 Ge10.1 55 −5.6
Fe86.6 Ge13.4 11 −1.0
Fe81.5 Ge18.5 −96 −5.25
Fe95 Si5 33 −4.3
Fe87.8 Si12.2 −10 1.1
Fe80.2 Si19.8 −21 2.3
Fe93.3 Al6.7 38 −4.7
Fe86.6 Al13.4 75 −8.11
Fe79.8 Al20.2 105 −8.31

Fig. 13 Contour plots of Fe-Al-Ga λ100 Fe-Ge-Ga


λ100 for Fe-Al-Ga (left) and
Fe-Ge-Ga (right) alloys. The Al +270 Ge
grid spacing is 5 atomic %.
Note the rapid change of sign +200
of λ100 with slight additions +120
of Ge in Fe-Ge-Ga. (Adapted
from [91]) +10
−27
−100
Fe Ga Fe Ga

The decrease of λ100 with increasing Ge content, leading to negative λ100 above
14% Ge content [91], has been tackled by theory [120], which finds that at small
Ge content a Ge-induced softening of the tetragonal shear modulus contributes to a
large magnetostriction. The calculated softening of the tetragonal shear modulus is
in qualitative agreement with experimental results [91]. At higher Ge content, the
availability of eg holes above the Fermi energy induces a negative magnetostriction
[120]. It has been predicted that the addition of Cu (Fe79.7 Ga18.7 Cu1.6 ) should lead
to an enhanced magnetostriction of λ100 = 550 × 10−6 [116], but this is not
observed. However, doping melt-spun Fe83 Ga17 with 0.23 atomic % Tb has been
reported to lead to a four times increased magnetostriction as compared to the
binary alloy [144]. This increase has been ascribed to a localized increased spin-
orbit interaction in proximity to Tb [144] (Fig. 13).

Zero Magnetostriction Alloys and Soft Magnetic Materials

Specific applications require ultra-soft magnetic materials. They are characterized


by small coercivity (low magnetization reversal fields) and high permeability.
Examples are permalloy (FeNi alloy with more than 35% Ni); mumetal (NiFe
11 Magnetostriction and Magnetoelasticity 581

alloy with roughly 80% Ni and small additions of other metals such as Cu, Cr,
Mo, balance Fe); sendust (Fe alloy with 10 weight % Si and 5% Al); Fe-Si
alloys, MnZn and NiZn ferrites, and amorphous (FeCo)B alloys; and nanocrystalline
Fe-based alloys, which find widespread applications in magnetic sensors, trans-
formers, electrical motors, and magnetic shielding [145, 146, 10]. Achieving
these material properties is innately linked with the demand for vanishingly
small magnetic anisotropy. An important contribution to magnetic anisotropy is
magnetoelastic coupling, and consequently small magnetostriction is required.
This reduces the strain-induced magnetic anisotropy, which arises inevitably in
manufacturing and shaping (bending) the material for use in applications. Thermal
annealing after manufacturing, often in an H2 atmosphere, may relax strain and is
often performed to obtain ultra-soft magnets.
Amorphous and nanocrystalline materials can be of high interest in correspond-
ing applications, as their effective magnetocrystalline anisotropy may average out
to zero, provided that the crystallographic grain size is smaller than the magnetic
domain size. One example is the alloy Fe73.5 Cu1 Nb3 Si13.5 B9 with a suppressed
magnetocrystalline anisotropy and zero magnetostriction [147].
In the quest for low magnetic anisotropy, one can exploit that different elements
contribute to magnetic anisotropy with opposing signs, leading to zero net magnetic
anisotropy. This venue is followed for permalloy, where the composition Ni78 Fe22
shows almost zero magnetocrystalline anisotropy and almost zero magnetostriction.
Bringing both properties to zero at exactly the same composition can be achieved
by adding further elements such as Mo and Cu. This is the material of choice when
highest initial permeability of up to 106 μ0 is required [10].
A critical aspect for application is that permalloy films show a magnetostrictive
behavior which depends on film thickness [56]. Figure 14 reveals that zero
magnetostriction is only observed for films thicker than 7 nm, whereas thinner films

Fig. 14 Room temperature


saturation magnetostriction of
sputtered polycrystalline 0.5
Ni81 Fe19 films, deposited on
amorphous alumina on a 0
Si(001) substrate. Films
thinner than 7 nm show -0.5
λS (10-6)

nonzero magnetostriction.
-1.0
(Adapted from [56])
-1.5

-2.0

-2.5

-3.0
0 5 10 15 20 25 30 35
thickness (nm)
582 D. Sander

exhibit magnetostriction of −2.5 × 10−6 at 3 nm. This suggests that other effects,
such as sample roughness, film strain, and electronic hybridization between film and
substrate, may lead to a qualitatively different magnetoelastic behavior in atomically
thin films as compared to bulk [17, 65, 50].

Magnetostriction Data for Paramagnetic Metals and Alloys

Experiments reveal for paramagnetic samples a magnetic field-induced strain λ


which is proportional to H 2 [90, 148]. The magnetostriction is ascribed to the
volume dependence of the magnetic susceptibility [149, 90, 150]. Table 7 gives the
slope of experimental data when λ is plotted as a function of H 2 .

Magnetostriction Data for Bi

Magnetostrictive strain measurements in Bi show pronounced oscillations, which


are ascribed to the de Haas-van Alphen effect [151]. At a field of 10 T, a longitudinal
magnetostriction of −10 × 10−6 is measured along the binary axis [151].

Magnetostriction Data for Tb, Dy, and Ho

Magnetostriction of the rare earth metals Tb, Dy, and Ho is large at cryogenic
temperature with strains of order 10−3 [152, 153]. These elements show different
magnetic order with increasing temperature, but none at room temperature.

Table 7 Experimental data Metal λ


of the magnetostriction in
Ti −0.69
10−10 /T2 along the field
direction obtained at liquid Zr −3.79
helium and liquid hydrogen V 4.18
temperatures in fields of up to Nb 3.6
10 T [90, 148] Ta 1.25
Mo 7.12
W 1.83
Ru −0.9
Rh 7.0
Pd −25
Ir 2.4
Pt −20.0
Rh50 Ir50 6.0
Rh50 Pd50 17.0
Pd67 Pt33 −11.0
Pd33 Pt67 −50
11 Magnetostriction and Magnetoelasticity 583

Magnetostriction measurements of Tb [152], Dy, and Ho [153] as a function of


field and at different temperatures reveal a considerable complexity, as indicated
in Fig. 15. This is ascribed to different magnetic phases, which are probed by
temperature and field variations [153]. Note the large magnitude of magnetostriction
λ of order 10−3 at low temperature, two–three orders of magnitude larger than 3d
transition metals (Table 2).
Comparably large magnetostrictive strains have been reported below 150 K and
in large fields of order several T for amorphous random anisotropy alloys containing
4f elements [154, 155].

(a) x10-3 Tb
3.2 78.7 K, λb
2.4
1.6 224.3 K, λc
0.8
224.5 K, λb
λ

0
79.7 K, λc
-0.8
c 225.4 K, λa
-1.6 a
-2.4 Hb
79.6 K, λa

0 0.4 0.8 1.2 1.6 2 2.4 2.8


H(T)
(b) x10-3 Dy
22 K, λa (c) x10-3 Ho
4
85 K, λa
3 144 K, λc 45 K, λc
3 77 K, λc
2 90 K, λc
2 4.2 K, λc
1 44 K, λc
1 4.2 K, λb
0
λ

c 20 K, λc 0 45 K, λb
λ

-1 a c
H a 77 K, λb
-2 b -1 77 K, λa
144 K, λb Hb 4.2 K, λa
-3 -2 45 K, λa
85 K, λb
-4 20 K, λb 0 0.5 1 1.5 2 2.5
0 0.5 1 1.5 2 2.5 H(T)
H(T)

Fig. 15 Magnetostrictive strain for (a) Tb, (b) Dy, and (c) Ho measured along different directions
with the field along the a and b direction for Dy and Tb and Ho, respectively. (Adapted from
[152, 153])
584 D. Sander

Magnetostriction Data of TbFe2 (Terfenol) and Tb27 Dy73 Fe2


(Terfenol-D)

The Laves phase alloys TbFe2 (Terfenol) and DyFe2 show a large magnetostriction
also at room temperature; see Fig. 16. Large magnetic fields are required to reach
saturation magnetostriction close to 1.3 × 10−3 of the isotropic polycrystal. The
necessity of high fields for saturation has been ascribed to the pronounced magnetic
anisotropy of the material. SmFe2 shows a comparable magnitude of magnetostric-
tion, although with opposite sign [21]. Curiously, the different magnetostrictive
coefficients differ vastly, λ111  λ100 , and crystals with predominant (111)
orientation give the largest strain [21].
Tb and Dy lead to fourth-order cubic magnetic anisotropy of opposite sign in
the respective rare earth-Fe2 compound. Thus, the alloy composition of Tb-Dy-
Fe can be tuned such as to minimize the magnetic anisotropy, still preserving a
large room temperature magnetostriction. The material has been named Terfenol-D
(Tb27 Dy73 Fe2 ). It reaches a saturation magnetostriction of order 2.2 × 10−3 at room
temperature in moderate fields of 0.3 T [21]. Alloying of Terfenol-D with other
elements has not lead to higher magnetostriction, but specific material properties
may be tuned by alloying [156].

x 10-3 x 10-3
1.0 295 K λll 3.5 TbFe2 (H= 8)
0.8
TbFe2
3.0
0.6
2.5 TbFe2 (2.5 T)
0.4 DyFe2
λll
0.2 2.0
3/2 λS
λ

0 1.5
-0.2 DyFe2
λ T
1.0
-0.4
TbFe2 0.5
-0.6 DyFe2 (2.5 T)
λ T
-0.8 0
0 1.0 2.0 3.0 0 50 100 150 200 250 300
H(T) T(K)

Fig. 16 Magnetostrictive strain at 295 K (left) and temperature dependence of magnetostriction


of the isotropic polycrystal 32 λS = λ − λ⊥ (right) for TbFe2 and DyFe2 . Note that a large
magnetostriction prevails at room temperature. (Adapted from [18])
11 Magnetostriction and Magnetoelasticity 585

Table 8 Magnetostriction Material λ100 λ111 λS


λ100 , λ111 , and λS (for
Fe3 O4 −19 81 41
polycrystals) in 10−6 at room
temperature, unless noted Fe3 O4 (124 K) −23 55 24
otherwise [128, 18, 11, 13] γ Fe2 O3 −5
CoFe2 O4 −670 120 −110
Co1.1 Fe1.9 O4 −250
Co0.9 Fe2.2 O4 −590 120 −210
Co0.3 Zn0.2 Fe2.2 O4 −210 110 −18
Co0.3 Mn0.4 Fe2.0 O4 −200 65 −40
Ni0.8 Fe2.2 O4 −36 −4 −17
MgFe2 O4 −6
Li0.5 Fe2.5 O4 −8
Mn0.98 Fe1.86 O4 −35 −1 −15
Mn0.6 Zn0.1 Fe2.1 O4 −14 14 3
NiFe2 O4 −26
Y3 Fe5 O12 (YIG) −1.4 −1.6 −2

Magnetostriction Data of Oxide Materials

Exchange interactions in oxides are usually antiferromagnetic, but they can lead to
ferromagnetic order in structures such as spinel or garnet. Some values are listed
in Table 8. Values for cobalt-containing are large on account of the large spin-orbit
interaction for the Co2+ cation.

Appendix

Notations for Lattice Strain

There are different conventions for the definition of strain, and it is mandatory
essential to distinguish between tensor strain ij and engineering strain γij [49,17].
The two conventions differ by a factor of two for the off-diagonal elements such
that 2ij = γij for i = j . Note that the use of the tensor strain is required if it
is transformed according to the tensor transformation rules to account for rotated
coordinate systems [49]. Throughout this contribution, all strain expressions are
given by the tensor strain. For completeness, we note that also the contracted Voigt
notation is often used. Starting from the tensor strain, the replacement rule given
in Table 9 applies to combine two subscripts into one. Note that factors of two
are inserted for the components with subscripts 4, 5, and6 to get 4 = 223 , 5 =
213 , and6 = 212 .
586 D. Sander

Table 9 Subscript replacement rule for the contracted Voigt notation. Note that factors of two are
inserted for the components with subscripts 4, 5, and6 to get 4 = 223 , 5 = 213 , and6 = 212
Convention Subscripts
Tensor notation: 11 22 33 23, 32 31, 13 12, 21
Contracted Voigt notation: 1 2 3 4 5 6

Relation Between λ and B for the Hexagonal System

We present here the relation between λi and Bj from Eq. (21).

B2 (c11 − c12 )c13 + B3 (−c11 + c12 )c33 + B1 (c13


2 −c c )
11 33
λA = (37)
a
B2 (c11 − c12 )c13 + B3 (−c11 + c12 )c33 + B1 (−c13
2 +c c )
11 33
λB = (38)
a
−B2 (c11 + c12 ) + (B1 + 2B3 )c13
λC = (39)
a
−B4 a + (−B2 (c11 − c12 )(c11 + c12 − c13 ) B3 (c11 − c12 )(2c13 − c33 )
λD = +
4ac44 4ac44
B1 (c11 c13 − c12 c13 + c13
2 − c c ))c
11 33 44
+ (40)
4ac44
a = (c11 − c12 )(−2c13
2
+ (c11 + c12 )c13 )

References
1. Becker, R., Döring, W.: Ferromagnetismus. Springer, Berlin (1939)
2. Kittel, C.: Physical theory of ferromagnetic domains. Rev. Mod. Phys. 21, 541–583 (1949)
3. Lee, E.W.: Magnetostriction and magnetomechanical effects. Rep. Prog. Phys. 18(1), 184
(1955)
4. Kneller, E.: Ferromagnetismus. Springer, Berlin (1962)
5. Kittel, C.: Introduction to Solid State Physics. Wiley, New York (1963)
6. Bozorth, R.M.: Ferromagnetism. IEEE-Press, Piscataway (1993)
7. Chikazumi, S.: Physics of Ferromagnetism. Oxford University Press, Oxford (2010)
8. du Trémolet de Lacheisserie, E.: Magnetostriction: Theory and Applications of Magnetoelas-
ticity. CRC-Press, Boca Raton (1993)
9. Hubert, A., Schäfer, R.: Magnetic Domains. Springer, Berlin (1998)
10. O’Handley, R.C.: Modern Magnetic Materials: Principles and Applications. Wiley, New York
(2000)
11. O’Handley, R.C.: Magnetostrictive materials and magnetic shape memory materials. In:
Kronmüller, H., Parkin, S. (eds.) Handbook of Magnetism and Advanced Magnetic Materials,
vol. 4, pp. 2401–2427. Wiley, New York (2007)
12. Moral, A.: Handbook of Magnetostriction and Magnetostrictive Materials. 2 volumes. Del
Moral Publisher, University of Zaragoza, Zaragoza (2008)
11 Magnetostriction and Magnetoelasticity 587

13. Coey, J.M.D.: Magnetism and Magnetic Materials. Cambridge University Press, Cambridge
(2010)
14. Joule, J.P.: On a new class of magnetic forces. Ann. Electricity Magn. Chem. 8, 219–224
(1842)
15. Joule, J.P.: On the effect of magnetism upon the dimensions of iron and steel bars. Philos.
Mag. (Third Series) 30, 76–87 (1847)
16. Joule, J.P.: On the effect of magnetism upon the dimensions of iron and steel bars. Philos.
Mag.(Third Series) 30, 225–241 (1847)
17. Sander, D.: The correlation between mechanical stress and magnetic anisotropy in ultrathin
films. Rep. Prog. Phys. 62, 809–858 (1999)
18. Clark, A.E., Belson, H.S.: Giant room-temperature magnetostrictions in TbFe2 and DyFe2 .
Phys. Rev. B 5(9), 3642–3644 (1972)
19. Clark, A.E.: Magnetostrictive Rare Earth-Fe2 Compounds. Ferromagnetic Materials, vol. 1.
Nort-Holland Publishing Company, Amsterdam (1980)
20. Clark, A.E., Restorff, J.B., Wun-Fogle, M., Lograsso, T.A., Schlagel, D.L.: Magnetostrictive
properties of body-centered cubic Fe-Ga and Fe-Ga-Al alloys. Magn. IEEE Trans. 36(5),
3238–3240 (2000)
21. Clark, A.E., Wun-Fogle, M.: Modern magnetostrictive materials: classical and nonclassical
alloys. Proc. SPIE 4699, 421–436 (2002)
22. Engdahl, G.: Chapter 1 – Physics of Giant magnetostriction. In: Engdahl, G. (ed.) Handbook
of Giant Magnetostrictive Materials. Electromagnetism, pp. 1–125. Academic, San Diego
(2000)
23. Wassermann, E.F.: Invar: moment-volume instabilities in transition metals and alloys. In:
Buschow, K.H.J., Wohlfarth, E.P. (eds.) Ferromagnetic Materials, vol. 5, pp. 237–322. Nort-
Holland Publishing Company, Amsterdam (1990)
24. Hausch, G.: Magnetovolume effects in invar alloys: spontaneous and forced volume magne-
tostriction. Phys. Status Solidi (A) 18(2), 735–740 (1973)
25. Wassermann, E.F.: The Invar problem. J. Magn. Magn. Mater. 100(1–3), 346–362 (1991)
26. Takenaka, K., Takagi, H.: Magnetovolume effect and negative thermal expansion in MnCu-
GeN. Mater. Trans. 47, 471–474 (2006)
27. Nagaoka, H., Honda, K.: Sur l’aimantation et la magnétostriction des aciers au nickel. J. Phys.
Theor. Appl. 3(1), 613–620 (1904)
28. Khmelevskyi, S., Mohn, P.: Magnetostriction in Fe-based alloys and the origin of the Invar
anomaly. Phys. Rev. B 69, 140404 (2004)
29. Shiga, M.: Invar alloys. Curr. Opin. Solid State Mater. Sci. 1(3), 340–348 (1996)
30. Khmelevskyi, S., Turek, I., Mohn, P.: Large negative magnetic contribution to the thermal
expansion in iron-platinum alloys: quantitative theory of the Invar effect. Phys. Rev. Lett. 91,
037201 (2003)
31. Villari, E.: Intorno alle modificazioni del momento magnetico di una verga di ferro e di
acciaio, prodotte per la trazione della medesima e pel passaggio di una corrente attraverso
la stessa. Il Nuovo Cimento 20(1), 317–362 (1864)
32. Siegel, S., Quimby, S.L.: The variation of Young’s modulus with magnetization and temper-
ature in nickel. Phys. Rev. 49, 663–670 (1936)
33. Hankemeier, S., Frömter, R., Mikuszeit, N., Stickler, D., Stillrich, H., Pütter, S., Vedmedenko,
E.Y., Oepen, H.P.: Magnetic ground state of single and coupled permalloy rectangles. Phys.
Rev. Lett. 103, 147204 (2009)
34. Berry, B.S., Pritchet, W.C.: Magnetic annealing and directional ordering of an amorphous
ferromagnetic alloy. Phys. Rev. Lett. 34, 1022–1025 (1975)
35. Becker, R., Kornetzki, M.: Einige magneto-elastische Torsionsversuche. Zeitschrift für Physik
88(9–10), 634–646 (1934)
36. Hathaway, K.B., Clark, A.E., Teter, J.P.: Magnetomechanical damping in giant magnetostric-
tion alloys. Metall. Mater. Trans. A 26(11), 2797–2801 (1995)
37. Wiedemann, G.: Über die Torsion und die Beziehung derselben zum Magnetismus. Ann.
Phys. 182, 161–201 (1859)
588 D. Sander

38. Drosdziok, S., Wessel, K.: A method for ultrasensitive magnetostriction measurement. Magn.
IEEE Trans. 9(1), 56–59 (1973)
39. Schmoller, F.v.: Untersuchungen über den Matteucci-Effekt. Zeitschrift für Physik 93(1–2),
35–51 (1935)
40. Dimitropoulos, P.D., Avaritsiotis, J.N.: A micro-fluxgate sensor based on the Matteucci effect
of amorphous magnetic fibers. Sensors Actuators A: Phys. 94(3), 165–176 (2001)
41. Szymczak, H.: From almost zero magnetostriction to giant magnetostrictive effects: recent
results. J. Magn. Magn. Mater. 200(1–3), 425–438 (1999)
42. Ullakko, K., Huang, J.K., Kantner, C., O’Handley, R.C., Kokorin, V.V.: Large magnetic-field-
induced strains in Ni2 MnGa single crystals. Appl. Phys. Lett. 69(13), 1966–1968 (1996)
43. O’Handley, R.C.: Model for strain and magnetization in magnetic shape-memory alloys. J.
Appl. Phys. 83(6), 3263–3270 (1998)
44. Sozinov, A., Likhachev, A.A., Lanska, N., Ullakko, K.: Giant magnetic-field-induced strain
in NiMnGa seven-layered martensitic phase. Appl. Phys. Lett. 80(10), 1746–1748 (2002)
45. Ye, M., Kimura, A., Miura, Y., Shirai, M., Cui, Y.T., Shimada, K., Namatame, H., Taniguchi,
M., Ueda, S., Kobayashi, K., Kainuma, R., Shishido, T., Fukushima, K., Kanomata, T.: Role
of electronic structure in the martensitic phase transition of Ni2 Mn1+x Sn1−x studied by hard-
X-ray photoelectron spectroscopy and ab initio calculation. Phys. Rev. Lett. 104, 176401
(2010)
46. Ibarra, M.R., Mahendiran, R., Marquina, C., García-Landa, B., Blasco, J.: Huge anisotropic
magnetostriction in La1−x Srx CoO3−δ (x >∼ 0.3) Field-induced orbital instability. Phys.
Rev. B 57, R3217–R3220 (1998)
47. Sato, K., Bartashevich, M.I., Goto, T., Kobayashi, Y., Suzuki, M., Asai, K., Matsuo, A.,
Kindo, K.: High-field magnetostriction of the spin-state transition compound LaCoO3 . J.
Phys. Soc. Jpn. 77(2), 024601 (2008)
48. Ikuta, H., Hirota, N., Nakayama, Y., Kishio, K., Kitazawa, K.: Giant magnetostriction in
Bi2 Sr2 CaCu2 O8 single crystal in the superconducting state and its mechanism. Phys. Rev.
Lett. 70, 2166–2169 (1993)
49. Nye, J.F.: Physical Properties of Crystals. Oxford University Press, Oxford (1985)
50. Sander, D., Kirschner, J.: Non-linear magnetoelastic coupling in monolayers: experimental
challenges and theoretical insights. Phys. Status Solidi (B) 248(10), 2389–2397 (2011)
51. Callen, E.R., Callen, H.B.: Static magnetoelastic coupling in cubic crystals. Phys. Rev. 129,
578–593 (1963)
52. Callen, E., Callen, H.B.: Magnetostriction, forced magnetostriction, and anomalous thermal
expansion in ferromagnets. Phys. Rev. 139, A455–A471 (1965)
53. O’Handley, R.C., Sun, S.-W.: Strained layers and magnetoelastic coupling. J. Magn. Magn.
Mater. 104–107(Part 3), 1717–1720 (1992)
54. O’Handley, R.C., Song, O.-S., Ballentine, C.A.: Determining thin-film magnetoelastic con-
stants. J. Appl. Phys. 74, 6302–6307 (1993)
55. Koch, R., Weber, M., Rieder, K.H.: Magnetoelastic coupling of Fe at high stress investigated
by means of epitaxial Fe(001) films. J. Magn. Magn. Mater. 159, L11–L16 (1996)
56. Kim, Y.K., Silva, T.J.: Magnetostriction characteristics of ultrathin permalloy films. Appl.
Phys. Lett. 68(20), 2885–2886 (1996)
57. Wedler, G., Walz, J., Greuer, A., Koch, R.: Stress dependence of the magnetoelastic coupling
constants B1 and B2 of epitaxial Fe(001). Phys. Rev. B 60(16), R11313–R11316 (1999)
58. Tian, Z., Sander, D., Kirschner, J.: Nonlinear magnetoelastic coupling of epitaxial layers of
Fe, Co, and Ni on Ir(100). Phys. Rev. B 79, 024432 (2009)
59. Akulov, N.S.: Über die Anwendungen des Gesetzes ferromagnetischer Anisotropie zur
Berechnung der Eigenschaften polykristallinischen Eisens. Zeitschrift für Physik 66(7–8),
533–542 (1930)
60. Callen, H.B., Goldberg, N.: Magnetostriction of polycrystalline aggregates. J. Appl. Phys.
36(3), 976–977 (1965)
61. Mason, W.P.: Derivation of magnetostriction and anisotropic energies for hexagonal, tetrago-
nal, and orthorhombic crystals. Phys. Rev. 96(2), 302–310 (1954)
11 Magnetostriction and Magnetoelasticity 589

62. Hubert, A., Unger, W., Kranz, J.: Messung der Magnetostriktionskonstanten des Kobalts als
Funktion der Temperatur. Zeitschrift für Physik 224(1–3), 148–155 (1969)
63. Wu, R., Freeman, A.J.: First principles determination of magnetostriction in transition metals.
J. Appl. Phys. 79, 6209–6212 (1996)
64. Hjortstam, O., Baberschke, K., Wills, J.M., Johansson, B., Eriksson, O.: Magnetic anisotropy
and magnetostriction in tetragonal and cubic Ni. Phys. Rev. B 55(22), 15026–15032
(1997)
65. Sander, D.: The magnetic anisotropy and spin reorientation of nanostructures and nanoscale
films. J. Phys.: Condens. Matter 16(20), R603 (2004)
66. Sander, D., Pan, W., Ouazi, S., Kirschner, J., Meyer, W., Krause, M., Müller, S., Hammer,
L., Heinz, K.: Reversible H-induced switching of the magnetic easy axis in Ni/Cu(001) thin
films. Phys. Rev. Lett. 93(24), 247203 (2004)
67. Marcus, P.M., Jona, F.: The strain tensor in a general pseudomorphic epitaxial film. J. Phys.
Chem. Solids 55, 1513–1519 (1994)
68. Marcus, P.M., Jona, F.: Strains in epitaxial films: the general case. Phys. Rev. B 51(8), 5263–
5268 (1995)
69. Jona, F., Marcus, P.M.: The importance of strain analysis in studies of pseudomorphic growth.
Surf. Rev. Lett. 4, 817–820 (1997)
70. Fähnle, M., Komelj, M.: Nonlinear magnetoelastic coupling coefficients from simultaneous
measurements of magnetostrictive stress and anisotropy in epitaxial films. J. Magn. Magn.
Mater. 220, L13–L17 (2000)
71. Komelj, M., Fähnle, M.: Nonlinear magnetelastic coupling coefficients in Fe from first
principles. J. Magn. Magn. Mater. 220, L8–L12 (2000)
72. Fähnle, M., Komelj, M.: Second-order magnetoelastic effects: from the Dirac equation to the
magnetic properties of ultrathin epitaxial films for magnetic thin-film applications. Zeitschrift
für Metallkunde/Mater. Res. Adv. Tech. 93(10), 970–973 (2002)
73. Fähnle, M., Komelj, M., Wu, R.Q., Guo, G.Y.: Magnetoelasticity of Fe: possible failure of
ab initio electron theory with the local-spin-density approximation and with the generalized-
gradient approximation. Phys. Rev. B 65(14), 144436 (2002)
74. Komelj, M., Fähnle, M.: Shear-strain-related nonlinear magnetoelastic properties of epitaxial
films. Phys. Rev. B 65(9), 092403 (2002)
75. Komelj, M., Fähnle, M.: Determination of the complete set of second-order magnetoelastic
coupling constants on epitaxial films. Phys. Rev. B 65(21), 212410 (2002)
76. Komelj, M., Fähnle, M.: On the magnetoelastic contribution to the magnetic anisotropy of
thin epitaxial Permalloy films: an ab initio study, J. Magn. Magn. Mater. 238(2–3), 125–128
(2002)
77. Komelj, M., Fähnle, M.: Nonlinear magnetoelastic behavior of the bcc phases of Co and Ni:
importance of third-order contributions for bcc Ni. Phys. Rev. B 73(1), 012404 (2006)
78. Ekreem, N.B., Olabi, A.G., Prescott, T., Rafferty, A., Hashmi, M.S.J.: An overview of
magnetostriction, its use and methods to measure these properties. J. Mater. Process. Technol.
191(1–3), 96–101 (2007)
79. Brizzolara, R.A., Colton, R.J.: Magnetostriction measurement using a tunneling-tip. J. Magn.
Magn. Mater. 88, 343–350 (1990)
80. Costa, J.L., Nogués, J., Rao, K.V.: Direct measurement of magnetostrictive process in
amorphous wires using a scanning tunneling microscope. J. Appl. Phys. 76, 7030 (1994)
81. White, G.K.: Measurement of thermal expansion at low temperatures. Cryogenics 1(3), 151–
158 (1961)
82. Brändli, G., Griessen, R.: Two capacitance dilatometers. Cryogenics 13(5), 299–302 (1973)
83. Tsuya, N., Arai, K.I., Ohmori, K., Shiraga, Y.: Magnetostriction measurement by three
terminal capacitance method. Jpn. J. Appl. Phys. 13(11), 1808 (1974)
84. Pott, R., Schefzyk, R.: Apparatus for measuring the thermal expansion of solids between 1.5
and 380 K. J. Phys. E: Sci. Instrum. 16(5), 444 (1983)
85. Hüller, K., Sydow, M., Dietz, G.: Magnetic anisotropy, magnetostriction and intermediate
range order in Co-P alloys. J. Magn. Magn. Mater. 53(3), 269–274 (1985)
590 D. Sander

86. Weber, M., Koch, R., Rieder, K.H.: UHV cantilever beam technique for quantitative mea-
surements of magnetization, magnetostriction, and intrinsic stress of ultrathin magnetic films.
Phys. Rev. Lett. 73(8), 1166–1169 (1994)
87. Rotter, M., Müller, H., Gratz, E., Doerr, M., Loewenhaupt, M.: A miniature capacitance
dilatometer for thermal expansion and magnetostriction. Rev. Sci. Instrum. 69(7), 2742–2746
(1998)
88. Schmiedeshoff, G.M., Lounsbury, A.W., Luna, D.J., Tracy, S.J., Schramm, A.J., Tozer, S.W.,
Correa, V.F., Hannahs, S.T., Murphy, T.P., Palm, E.C., Lacerda, A.H., Bud’ko, S.L., Canfield,
P.C., Smith, J.L., Lashley, J.C., Cooley, J.C.: Versatile and compact capacitive dilatometer.
Rev. Sci. Instrum. 77(12), 123907 (2006)
89. Küchler, R., Bauer, T., Brando, M., Steglich, F.: A compact and miniaturized high resolution
capacitance dilatometer for measuring thermal expansion and magnetostriction. Rev. Sci.
Instrum. 83(9), 095102 (2012)
90. Fawcett, E.: Magnetostriction of paramagnetic transition metals. I. Group 4 – Ti and Zr; Group
5 – V, Nb, and Ta; Group 6 – Mo and W. Phys. Rev. B 2(6), 1604–1613 (1970)
91. Restorff, J.B., Wun-Fogle, M., Hathaway, K.B., Clark, A.E., Lograsso, T.A., Petculescu, G.:
Tetragonal magnetostriction and magnetoelastic coupling in Fe-Al, Fe-Ga, Fe-Ge, Fe-Si, Fe-
Ga-Al, and Fe-Ga-Ge alloys. J. Appl. Phys. 111(2)023905, 1–12 (2012)
92. Liniers, M., Madurga, V., Vázquez, M., Hernando, A.: Magnetostrictive torsional strain in
transverse-field-annealed Metglas ő 2605. Phys. Rev. B 31(7), 4425–4432 (1985)
93. Mitra, A., Vázquez, M.: Measurement of the saturation magnetostriction constant of amor-
phous wires, J. Appl. Phys. 67, 4986–4988 (1990)
94. Yamasaki, J., Ohkubo, Y., Humphrey, F.B.: Magnetostriction measurement of amorphous
wires by means os small-angle magnetization rotation. J. Appl. Phys. 67, 5472–5474
(1990)
95. Eastman, D.E.: Magneto-elastic coupling in RbMnF3 . Phys. Rev. 156(2), 645–654 (1967)
96. Smith, A.B.: A ferromagnetic resonance method for measuring magnetostriction constants.
Rev. Sci. Instrum. 39, 378–385 (1967)
97. Krams, P., Hillebrands, B., Güntherodt, G., Oepen, H.P.: Magnetic anisotropies of ultrathin
Co films on Cu(1 1 13) substrates. Phys. Rev. B 49(5), 3633–3636 (1994)
98. Premper, J., Sander, D., Kirschner, J.: A combined surface stress and magneto-optical Kerr
effect measurement setup for temperatures down to 30 K and in fields of up to 0.7 T. Rev. Sci.
Instrum. 83(7), 073904 (2012)
99. Dahmen, K., Ibach, H., Sander, D.: A finite element analysis of the bending of crystalline
plates due to anisotropic surface and film stress applied to magnetoelasticity. J. Magn. Magn.
Mater. 231(1), 74–84 (2001)
100. Dahmen, K., Lehwald, S., Ibach, H.: Bending of crystalline plates under the influence of
surface stress — a finite element analysis. Surf. Sci. 446(1–2), 161–173 (2000)
101. Klokholm, E.: The measurement of magnetostriction in ferromagnetic thin films. IEEE Trans.
Magn. MAG-12, 819–821 (1976)
102. Tam, A.C., Schroeder, H.: Precise measurement of a magnetostriction coefficient of a thin
sofmagnetic film deposited on a substrate. J. Appl. Phys. 64, 5422–5424 (1988)
103. Betz, J., du Trémolet de Lacheisserie, E., Baczewski, L.T.: Magnetoelastic properties of nickel
thin films. Appl. Phys. Lett. 68, 132–133 (1996)
104. Koch, R., Weber, M., Henze, E., Rieder. K.H.: Magnetization, magnetostriction and intrinsic
stress of polycrystalline Fe films measured by a UHV cantilever beam technique. Surf. Sci.
331–333, 1398–1403 (1995)
105. Wu, R.: First principles determination of magnetic anisotropy and magnetostriction in transi-
tion metal alloys. In: Baberschke, K., Nolting, W., Donath, M. (eds.) Band-Ferromagnetism:
Ground-State and Finite-Temperature Phenomena, pp. 60–71. Springer, Berlin/Heidelberg
(2001)
106. Wu, R.: Theory of magnetocrystalline aniostropy and magnetoelasticity in transition metal
system. In: Kronmüller, H., Parkin, S. (eds.) Handbook of Magnetism, vol. 1. Wiley, New
York (2007)
11 Magnetostriction and Magnetoelasticity 591

107. Chappert, C., Bruno, P.: Magnetic anisotropy in metallic ultrathin films and related experi-
ments on cobalt. J. Appl. Phys. 64, 5736–5741 (1988)
108. Bruno, P., Renard, J.-P.: Magnetic surface anisotropy of transition metal ultrathin films. Appl.
Phys. A 49, 499–506 (1989)
109. de Jonge, W.J.M., Bloemen, P.J.H., den Broeder, F.J.A.: Experimental investigation of
magnetic anisotropy. In: Bland, J., Heinrich, B. (eds.) Ultrathin Magnetic Structures I, pp.
65–90. Springer, Berlin (1994)
110. Johnson, M.T., Bloemen, P.J.H., den Broeder, F.J.A., de Vries, J.J.: Magnetic anisotropy in
metallic multilayers. Rep. Prog. Phys. 59, 1409–1458 (1996)
111. Skomski, R., Coey, J.M.D.: Permanent Magnetism. IOP Publishing, Bristol (1999)
112. Baberschke, K.: Anisotropy in magnetism. In: Baberschke, K., Nolting, W., Donath, M.
(eds.) Band-Ferromagnetism: Ground-State and Finite-Temperature Phenomena, pp. 27–45.
Springer, Berlin/Heidelberg (2001)
113. Burkert, T., Eriksson, O., James, P., Simak, S.I., Johansson, B., Nordström, L.: Calculation of
uniaxial magnetic anisotropy energy of tetragonal and trigonal Fe, Co, Ni. Phys. Rev. B 69,
104426–1–104426–7 (2004)
114. Blügel, S., Bihlmayer, G.: Magnetism of low-dimensional systems: theory. In: Kronmüller,
H., Parkin, S. (eds.) Handbook of Magnetism, vol. 1, pp. 598–639. Wiley, New York (2007)
115. Wu, R., Chen, L.J., Shick, A., Freeman, A.J.: First-principles determinations of magneto-
crystalline anisotropy and magnetostriction in bulk and thin-film transition metals. J. Magn.
Magn. Mater. 177–181, 1216–1219 (1998)
116. Wang, H., Zhang, Y.N., Wu, R.Q., Sun, L.Z., Xu, D.S., Zhang, Z.D.: Understanding strong
magnetostriction in Fe100−x Gax alloys. Sci. Rep. 3, 3521 (2013)
117. Bruno, P.: Tight-binding approach to the orbital magnetic moment and magnetocrystalline
anisotropy of transition-metal monolayers. Phys. Rev. B 39(1), 865–868 (1989)
118. Weinberger, P.: Magnetic Anisotropies in Nanostructured Materials. CRC, Boca Raton
(2008)
119. Wu, R., Chen, L., Freeman, A.J.: First principles determination of magnetostriction in bulk
transition metals and thin films. J. Magn. Magn. Mater. 170(1–2), 103–109 (1997)
120. Cao, J.X., Zhang, Y.N., Ouyang, W.J., Wu, R.Q.: Large magnetostriction of Fe(1−x) Gex and
its electronic origin: density functional study. Phys. Rev. B 80, 104414 (2009)
121. Cao, H., Gehring, P.M., Devreugd, C.P., Rodriguez, Rodriguez-Rivera, J.A., Li, J., Viehland,
D.: Role of nanoscale precipitates on the enhanced magnetostriction of heat-treated galfenol
(Fe1−x Gax ) alloys. Phys. Rev. Lett. 102, 127201 (2009)
122. Zhang, Y.N., Cao, J.X., Barsukov, I., Lindner, J., Krumme, B., Wende, H., Wu, R.Q.:
Magnetocrystalline anisotropy of Fe-Si alloys on MgO(001). Phys. Rev. B 81, 144418
(2010)
123. Buck, S., Fähnle, M.: Ab initio calculation of the giant magnetostriction in terbium and
erbium. Phys. Rev. B 57, R14044–R14047 (1998)
124. Hummler, K., Fähnle, M.: Full-potential linear-muffin-tin-orbital calculations of the magnetic
properties of rare-earth–transition-metal intermetallics. I. Description of the formalism and
application to the series R Co5 ( R =rare-earth atom). Phys. Rev. B 53, 3272–3289 (1996)
125. Kanamori, J.: Theory of the magnetic properties of ferrous and cobaltous oxides, I. Prog.
Theor. Phys. 17(2), 177–196 (1957)
126. Slonczewski, J.C.: Anisotropy and magnetostriction in magnetic oxides. J. Appl. Phys. 32(3),
S253–S263 (1961)
127. Greenwald, S.: The antiferromagnetic structure deformations in CoO and MnTe. Acta
Crystallogr. 6(5), 396–398 (1953)
128. Bozorth, R.M., Tilden, E.F., Williams, A.J.: Anisotropy and magnetostriction of some ferrites.
Phys. Rev. 99, 1788–1798 (1955)
129. Sander, D., Phark, S.-H., Corbetta, M., Fischer, J.A., Oka, H., Kirschner, J.: The impact
of structural relaxation on spin polarization and magnetization reversal of individual nano
structures studied by spin-polarized scanning tunneling microscopy. J. Phys.: Condens. Matter
26(39), 394008 (2014)
592 D. Sander

130. Brovko, O.O., Bazhanov, D.I., Meyerheim, H.L., Sander, D., Stepanyuk, V.S., Kirschner, J.:
Effect of mesoscopic misfit on growth, morphology, electronic properties and magnetism of
nanostructures at metallic surfaces. Surf. Sci. Rep. 69(4), 159–195 (2014)
131. Oka, H., Ignatiev, P.A., Wedekind, S., Rodary, G., Niebergall, L., Stepanyuk, V.S., Sander, D.,
Kirschner, J.: Spin-dependent quantum interference within a single magnetic nanostructure.
Science 327(5967), 843–846 (2010)
132. Oka, H., Brovko, O.O., Corbetta, M., Stepanyuk, V.S., Sander, D., Kirschner, J.: Spin-
polarized quantum confinement in nanostructures: scanning tunneling microscopy. Rev. Mod.
Phys. 86, 1127–1168 (2014)
133. Wu, R., Freeman, A.J.: Spin-orbit induced magnetic phenomena in bulk metals and their
surfaces and interfaces. J. Magn. Magn. Mater. 200(1–3), 498–514 (1999)
134. Gutjahr-Löser, T., Sander, D., Kirschner, J.: Magnetoelastic coupling in Ni and Fe monolayers
on Cu(001). J. Appl. Phys. 87, 5920–5922 (2000)
135. Gutjahr-Löser, T., Sander, D., Kirschner, J.: Magnetoelastic coupling in Co thin films on
W(001). J. Magn. Magn. Mater. 220, L1–L7 (2000)
136. Summers, E.M., Lograsso, T.A., Wun-Fogle, M.: Magnetostriction of binary and ternary
Fe–Ga alloys. J. Mater. Sci. 42(23), 9582–9594 (2007)
137. Clark, A.E., Hathaway, K.B., Wun-Fogle, M., Restorff, J.B., Lograsso, T.A., Keppens, V.M.,
Petculescu, G., Taylor, R.A.: Extraordinary magnetoelasticity and lattice softening in bcc Fe-
Ga alloys. J. Appl. Phys. 93(10), 8621–8623 (2003)
138. Wu, R.: Origin of large magnetostriction in FeGa alloys. J. Appl. Phys. 91(10), 7358–7360
(2002)
139. Wang, H., Zhang, Y.N., Yang, T., Zhang, Z.D., Sun, L.Z., Wu, R.Q.: Ab initio studies of
the effect of nanoclusters on magnetostriction of Fe1−x Gax alloys. Appl. Phys. Lett. 97(26),
262505 (2010)
140. Khachaturyan, A.G., Viehland, D.: Structurally heterogeneous model of extrinsic magne-
tostriction for Fe-Ga and similar magnetic alloys: part II. Giant magnetostriction and elastic
softening. Metall. Mater. Trans. A 38(13), 2317–2328 (2007)
141. Bhattacharyya, S., Jinschek, J.R., Khachaturyan, A., Cao, H., Li, J.F., Viehland, D.: Nanodis-
persed DO3 -phase nanostructures observed in magnetostrictive Fe-19 % Ga Galfenol alloys.
Phys. Rev. B 77, 104107 (2008)
142. Xing, Q., Lograsso, T.A.: Magnetic domains in magnetostrictive Fe–Ga alloys. Appl. Phys.
Lett. 93(18), 182501–1–182501–3 (2008)
143. Cao, H., Gehring, P.M., Devreugd, C.P., Rodriguez-Rivera, J.A., Li, J., Viehland, D.: Role of
nanoscale precipitates on the enhanced magnetostriction of heat-treated galfenol (Fe1−x Gax )
alloys. Phys. Rev. Lett. 102, 127201 (2009)
144. Wu, W., Liu, J., Jiang, C., Xu, H.: Giant magnetostriction in Tb-doped Fe83 Ga17 melt-spun
ribbons. Appl. Phys. Lett. 103(26), 262403–1–262403–3 (2013)
145. O’Handley, R.C.: Physics of ferromagnetic amorphous alloys. J. Appl. Phys. 62(10), R15–
R49 (1987)
146. Jiles: Magnetism and Magnetic Materials. Chapman&Hall, London (1994)
147. Herzer, G.: Modern soft magnets: amorphous and nanocrystalline materials. Acta Mater.
61(3), 718–734 (2013)
148. Fawcett, E.: Magnetostriction of paramagnetic transition metals. II. Group-VIII metals Ru,
Rh, Pd, Ir, Pt, and their alloys. Phys. Rev. B 2, 3887–3890 (1970)
149. Kapitza, P.: The study of the magnetic properties of matter in strong magnetic fields. Part V.
Experiments on magnetostriction in dia- and para-magnetic substances. Proc. R. Soc. Lond.
A: Math. Phys. Eng. Sci. 135(828), 568–600 (1932)
150. Chandrasekhar, B.S., Fawcett, E.: Magnetostriction in metals. Adv. Phys. 20(88), 775–794
(1971)
151. Michenaud, J.P., Heremans, J., Shayegan, M., Haumont, C.: Magnetostriction of bismuth in
quantizing magnetic fields. Phys. Rev. B 26, 2552–2559 (1982)
152. Rhyne, J.J., Legvold, S.: Magnetostriction of Tb single crystals. Phys. Rev. 138, A507–A514
(1965)
11 Magnetostriction and Magnetoelasticity 593

153. Legvold, S., Alstad, J., Rhyne, J.: Giant magnetostriction in dysprosium and holmium single
crystals. Phys. Rev. Lett. 10, 509–511 (1963)
154. del Moral, A., Arnaudas, J.I.: Magnetostriction of rare-earth random magnetic anisotropy spin
glasses. Phys. Rev. B 39, 9453–9466 (1989)
155. del Moral, A., Algarabel, P.A., Arnaudas, J.I., Benito, L., Ciria, M., de la Fuente, C., García-
Landa, B., Ibarra, M.R., Marquina, C., Morellón, L., de Teresa, J.M.: Magnetostriction effects.
J. Magn. Magn. Mater. 242–245, Part 2(0), 788–796 (2002)
156. Liu, J.H., Jiang, C.B., Xu, H.B.: Giant magnetostrictive materials. Sci. China Technol. Sci.
55(5), 1319–1326 (2012)

Dirk Sander received his PhD from RWTH Aachen in 1992 and
worked at the IBM T.J. Watson Research Center at Yorktown
Heights, New York, before he joined the Max Planck Institute of
Microstructure Physics, Halle, in 1993. He received his Habilita-
tion from Halle University in 2000. His research interests include
surface and film stress, magnetoelasticity, and spin-dependent
phenomena and superconductivity on the nanoscale.
Magnetoelectrics and Multiferroics
12
Jia-Mian Hu and Long-Qing Chen

Contents
Magnetoelectrics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 596
Multiferroics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 598
The Evolving Terminology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 598
Electric Field Switching of Magnetization in Multiferroic BiFeO3 :
Status and Perspective . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 602
Composite Multiferroics and Magnetoelectrics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 607
Terminology and Exiting Reviews . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 607
Mechanisms and Application of Magnetoelectric Effects and
Experimental Data of Magnetoelectric Coefficients in Composite Magnetoelectrics . . . . . 609
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 614

Abstract

Magnetoelectrics and multiferroics can possess mutually coupled magnetic and


ferroelectric order and thus have been utilized in exploring and designing many
novel multifunctional devices such as sensors, transducers, and memories. This
chapter presents a brief introduction to the terminology and classification of mag-
netoelectrics and multiferroics as well as the mechanisms underlying different
types of magnetoelectric couplings. Both single-phase and composite materials

J.-M. Hu ()
Department of Materials Science and Engineering, University of Wisconsin-Madison, Madison,
WI, USA
e-mail: jhu238@wisc.edu
L.-Q. Chen
Materials Research Institute, and Department of Materials Science and Engineering,
The Pennsylvania State University, University Park, PA, USA
e-mail: lqc3@psu.edu

© Springer Nature Switzerland AG 2021 595


J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_12
596 J.-M. Hu and L.-Q. Chen

are discussed. Experimental data showing the basic magnetic and ferroelectric
properties of different types of single-phase multiferroics are collected. For
composite magnetoelectrics, a relatively comprehensive experimental dataset of
both the direct and inverse magnetoelectric coefficients is compiled.

Magnetoelectrics

Magnetoelectrics are a family of solid-state materials that enables mutual conversion


between magnetic energy and electric energy in the absence of an electric current.
This is based on the coupling between the electric polarization P and a magnetic
field H, dP i = αijH dH j or between the magnetization M and an electric field E,
μ0 dM i = αijE dE j , where αijH and αijE are the direct and inverse magnetoelectric
coefficients, respectively, and μ0 is the vacuum permeability. In Landau theory,
the magnetoelectric effect in a single-phase material is typically described by
introducing an additional energy density (J/m3 ) term −α ij Ei Hj in the total free
energy of the system. The magnetoelectric effect can be understood by analogy
to other types of coupling effects in functional materials (Fig. 1). For example,
piezoelectrics enable mutual conversion between mechanical energy and electric
energy in the absence of an electric current, based on the piezoelectric effect –
the coupling between the mechanical strain ε and E or between P and the
mechanical stress σ . These coupling effects can find applications in devices such as
sensors, actuators, memories, and many other devices that rely on solid-state energy

Fig. 1 Coupling effects in


functional materials that
permit solid-state energy
conversion, showing order
parameters (blue) and
conjugate fields (yellow),
notably electric field E,
magnetic field H, and
temperature T. Other symbols
represent coupling
coefficients
12 Magnetoelectrics and Multiferroics 597

conversion. Furthermore, devices based on these coupling effects are generally


energy-efficient because the need for an electric current is obviated during energy
conversion.
The abovementioned P – H and M – E correlations were first proposed by
Curie in 1894 based on symmetry considerations [1] and then termed “magne-
toelectric” by Debye in 1926 [2]. However, in the 1920s–1950s, all attempts to
experimentally observe such effects had failed, and some scientists believed that
“no magnetoelectric effect can exist” (see a summary of these early works in a
text book by O’Dell [3]). In 1957, Landau and Lifshtiz first [4] pointed out that
the magnetoelectric effect is only allowed in time-asymmetric media (i.e., a time
reversal transformation t → − t changes the sign of thermodynamic potential of
a medium). Such broken time-reversal symmetry can be achieved extrinsically by
moving a dielectric medium [5, 6], applying a constant magnetic field to a paramag-
netic crystal [7], and appears naturally in a magnetically ordered (ferromagnetic,
ferromagnetic, and antiferromagnetic) crystal. In 1960, Dzyaloshinskii [8] first
pointed out that the magnetoelectric coupling should exist in the antiferromagnetic
Cr2 O3 , which was experimentally confirmed by Astrov [9] and then others [10–
12]. The magnetoelectric effect was soon found in about 80 different magnetically
ordered single-phase materials from the 1960s to the early 1970s. Examples include
Ti2 O3 [13], Gd2-x Fex O3 (a ferromagnet) [14], a few compounds of boracites (e.g.,
Ni3 B7 O13 I [15], a ferroelectric ferromagnet), phosphates (e.g., LiCoPO4 [16, 17]),
manganites (e.g., Ta2 Mn4 O9 [18]), and the PbFe0.5 Nb0.5 O3 solid solutions [19]. A
detailed summary of these early works can be found in an article by Schmid [20].
Also in the early 1970s, the concept of composite magnetoelectrics/multiferroics
and corresponding materials emerged, as will be discussed in Sect. “Composite
Multiferroics and Magnetoelectrics”.
The magnetoelectric effect is generally small in single-phase materials. For
instance, the peak value of αzzE in Cr O single crystal (the subscript z represents the
2 3
Cartesian axis) was found to be 4.13 × 10−12 s/m [11], which suggests that applying
an electric field of 106 V/m can only flip 4 or 5 spins out of one million spins
in the antiferromagnetic lattice [21]. In 1968, Brown (who is best known for his
contribution to the theory of micromagnetics [22]) and coworkers [23] proposed an
upper bound for the magnetoelectric coefficients α ij in single-phase materials, αij2 ≤
ε0 μ0 εii μjj , where ε0 is the vacuum permittivity; the diagonalized tensors εii and
μjj represent the relative permittivity and relative permeability, respectively. This
equation naturally leads to the following two conclusions. First, magnetoelectric
effects in ferroelectric (or ferromagnetic) materials that have large relative electric
permittivity (or magnetic permeability) should also be large. Second, if a material is
simultaneously ferroelectric and ferromagnetic, which is one type of “muliferroic”
material according to the definition by Schmid in his 1994 paper “Multi-ferroic
magnetoelectrics” [24], large magnetoelectric effects may emerge. Table 1 lists the
magnetoelectric coefficients of several single-phase magnetoelectrics.
598 J.-M. Hu and L.-Q. Chen

Table 1 Linear and nonlinear magnetoelectric coefficients in some single-phase magnetoelectrics


Linear
magneto- Testing
electric Testing tempera-
Materiala α ij (s/m)b effect? condition ture Reference
Cr2 O3 (bulk single 4.13 × 10−12 Yes M||Ec 265 K [11]
crystal)
Cr2 O3 (thin-film) (4.6 ± 0.3) × 10−12 Yes M||E 250 K [25]
TbPO4 (bulk single 3 × 10−10 Yes P||H 4.2 K [26]
crystal)
Ba0.52 Sr2.48 Co2 Fe24 O41 3.2 × 10−9 No P||H 305 K [27]
(bulk single crystal) @10.5mT
BiFeO3 (mono-domain 1.1 × 10−10 @ No P⊥H 4.2 K [28]
bulk single crystal) 24 Td
TbMnO3 (bulk single 1.1 × 10−9 @ 5 T No P⊥H 9K [29]
crystal)
TbMn2 O5 (bulk single 2 × 10−9 @ 1.1 T No P⊥H 3K [30]
crystal)
Ni3 B7 O13 I (bulk 1.6 × 10−9 Yese P⊥H 46 K [15]
single crystal)
a Bulk single crystal and thin-film materials are classified based on their method of growth
b The direct magnetoelectric coefficient αijH = dP i /dH j and the inverse magnetoelectric
coefficient αijE = μ0 dM i /dE j have the same unit of second/m in SI units; moreover, αijH = αijE in
the case of linear magnetoelectric effect. In the case of nonlinear magnetoelectric effect, αijH and
αijE would change with the driving magnetic and electric field, respectively, and we only list the
peak values of αijH for these materials along with the testing magnetic field, because the data of
nonlinear αijE are not available
c The direction that the induced M was recorded along is parallel to the applied electric field E
i j
(i = j), likewise for the polarization Pi induced by applied magnetic field Hj
d A large magnetic field is needed to break the spin cycloid such that the magnetoelectric effect

can become appreciable. The data measured on the same sample indicates αijH <3 × 10−11 when
H < 23 T
e Ni B O I can also display a nonlinear magnetoelectric effect: its spontaneous polarization
3 7 13
(0.076 μC cm−2 ) changes signs sharply at H ∼ ±0.6 T

Multiferroics

The Evolving Terminology

Multiferroics are a family of materials that “have two or more primary ferroic
properties in the same phase,” based on Schmid’s definition [24] (Fig. 2a). The
term ferroic was first introduced in 1970 by Aizu [35], who suggested that a
ferroic crystal should have two or more orientation states in the absence of a
magnetic field, electric field, or mechanical stress and can be switched from one
orientation state to another by a magnetic field, an electric field, a mechanical stress,
12 Magnetoelectrics and Multiferroics 599

Fig. 2 The evolving terminology of multiferroics. The original definition by Schmid [24] requires
the coexistence of two or all three primary ferroic order parameters, including spontaneous
magnetization Ms , spontaneous polarization Ps , and spontaneous strain εs , as shown in (a)
Eerenstein et al. [31] later suggested excluding the ferroelasticity. The present use of multiferroics
has been broadened to include antiferroic order (as stated in Refs. [32, 33]). (b) The Venn diagram,
drawn based on similar diagrams in Refs. [31, 34]. Some representative single-phase materials
are also listed. Magnetically (electrically) polarizable materials also include antiferromagnetic
(antiferroelectric) and other magnetically (electrically) ordered materials

or a combination of these stimuli. Ferroelectric, ferromagnetic, and ferroelastic (a


term also introduced by Aizu [36]) materials are classified as primary ferroics by
Newnham [37] in 1974, in which the primary order parameters are, respectively,
the spontaneous polarization Ps , spontaneous magnetization Ms , or spontaneous
strain εs .
Materials with the coexistence of Ps and Ms (ferroelectric ferromagnets) rep-
resent the most familiar example of multiferroics based on Schmid’s definition.
Ferromagnetic martensites such as NiMnGa, which simultaneously exhibit Ms
and εs , are also multiferroic. However, as pointed out by Eerenstein et al. [31]
in 2006, “the current trend is to exclude the requirement for ferroelasticity in
practice . . . .” According to the Venn diagram given in Ref. [31] (reproduced here in
Fig. 2b), multiferroics only refer to ferroelectric ferromagnets. The present use of
multiferroics has been broadened to include antiferroic (mainly the antiferroelectric
and antiferromagnetic) order [32, 33]. Notably, a ferroic order does not necessarily
need to be a primary order parameter in such broadened definition. For example,
improper ferroelectrics such as YMnO3 and TbMnO3 , where the polarization is
not the primary order parameter (i.e., ferroelectricity is induced geometrically or
by spin ordering, as will be discussed later), are often called as multiferroic in the
literature. In the remaining part of this chapter, we will use the broadened definition
for multiferroics. In the present section, we will briefly discuss the representative
materials of different types of single-phase multiferroics. Detailed information can
be found in a series of review articles [21, 31–33, 38–44].
We classify the single-phase multiferroics by the type of spin ordering, including
the ferroelectric ferromagnets, the ferroelectric ferrimagnets, and the ferroelectric
antiferromagnets (collinear and noncollinear). Moreover, Khomskii [45] classified
single-phase multiferroics into two types: Type I where the ferroelectricity and
magnetic order arise independently and Type II where the ferroelectric polarization
600 J.-M. Hu and L.-Q. Chen

is induced by magnetic order. Note that hexaferrites such as Ba0.5 Sr1.5 Zn2 Fe12 O22
[46] are not multiferroic, because they do not exhibit a spontaneous polarization
although a polarization on the order of 0.02 μC cm−2 can be induced and then
rotated by an applied magnetic field. Yet, these hexaferrites are magnetoelectric
because both magnetic field-controlled polarization [46–50] and electric field-
controlled magnetization [27, 51–54] have been experimentally demonstrated. More
information can be found in a topical review [55].

Ferroelectric Ferromagnets
Ferroelectric ferromagnets have remained rare in nature (notably in oxide per-
ovskites [56]), partly because ferromagnets are usually good electron conductors
while ferroelectrics are generally electron insulators. To our knowledge, in all
ferroelectric ferromagnets discovered so far, the ferroelectricity and magnetic order
arise independently (Type I multiferroics). For example, in the Ni3 B7 O13 I bulk
crystal, which was the first ferroelectric ferromagnet to be discovered [15], the
ferromagnetism arises from the transition metal Ni while the ferroelectricity arises
from the distorted I-O octahedron. In BiMnO3 (reported in both ceramics [57–
59] and thin films [58, 60]) and LaBiMnO3 thin films [61], ferroelectricity arises
from the 6 s lone pair of the Bi3+ , while the ferromagnetism arises from the Mn
cations. A special example is EuTiO3 : it is a paraelectric antiferromagnet in the form
of unstrained bulk crystal [62], yet can transform into the strongest ferroelectric
ferromagnet under a large (>1.2%) biaxial compressive strain in thin-film form [63,
64]. Table 2 summarizes the spontaneous/saturation polarization , magnetization,
and the onset temperatures for the ferroelectric order (TFE ) and ferromagnetic order
(TC ) of these ferroelectric ferromagnets.

Table 2 Properties of the ferroelectric-ferromagnet type single-phase multiferroics


Type I or
II multi-
Material ferroics Polarization Magnetization TFE TC Reference
Ni3 B7 O13 I Type I 0.076 μC 3.26 μB per Ni 60 K 64 K [15, 31, 65]
cm−2 @46 K atom @ 6 Ka
BiMnO3 Type I 0.15 μC 3.6 μB per Mn 450– 105 K [57, 58, 66,
cm−2 atom @5 K 490 K 67]
@87 Kb
LaBiMnO3 Type I N/Ac 2.5 × 105 A N/A 90 K [61]
m−1 @10 K
Strained Type I 10 μC cm−2 3 μB per Eu 250 K 4K [63]
EuTiO3 @5 Kd atom @ 1.8 Ke
a Saturating magnetic moment
b Saturating polarization
c Ferroelectricity demonstrated through piezoresponse force microscopy at 300 K
d Spontaneous polarization
e Spontaneous magnetic moment. μ : Bohr magneton
B
12 Magnetoelectrics and Multiferroics 601

Ferroelectric Antiferromagnets
Ferroelectric antiferromagnets are the most commonly reported single-phase mul-
tiferroics. These materials are grouped by their different origins of ferroelectricity
[32, 33], as follows:

• Solid solutions where the ferroelectric polarization arises from the 6 s lone pair of
the A-site atom (A = Bi, Pb) in the ABO3 structure that distorts the unit cell (viz.,
lone pair-driven ferroelectricity). Examples include Pb(Fe1/2 Nb1/2 )O3 that was
the first discovered material of this kind [68], and BiFeO3 (Sect. “Electric Field
Switching of Magnetization in Multiferroic BiFeO3 : Status and Perspective”),
the best known [69].
• Compounds where the ferroelectric polarization arises from the polar tilts and
rotations of the anionic sublattice (viz., geometrically driven ferroelectricity),
including hexagonal (h-) manganites RMnO3 (R = Sc, Y, In, or Dy-Lu, see, e.g.,
[70]), hexagonal LuFeO3 thin films [71], and some fluorides such as BaMF4
(M = Ni, Mn, Fe, Co, see, e.g., [72–74] and a review [75]).
• Compounds where the ferroelectricity polarization is induced by spin ordering
(viz., spin-driven ferroelectricity, see reviews [76–78]). Specifically, the polar-
ization can arise from:
(i) The noncollinear spin ordering through the inverse Dzyaloshinskii-Moriya
interaction (asymmetric exchange coupling) [79, 80]. Examples include
Cr2 BeO4 (the first material of this kind to be discovered) [81], orthorhombic
(o-) rare earth (RE) manganites REMnO3 with a cycloidal spin spiral
(Fig. 3a) (e.g., TbMnO3 [29]), and the CaMn7 O12 with a helical spin spiral
(Fig. 3b) [82, 83].
(ii) The collinear spin ordering through Heisenberg (symmetric) exchange
coupling [84, 85]. Examples include o-REMn2 O5 (e.g., TbMn2 O5 , the first
material of this kind to be discovered [30]), GdMn2 O5 [86], and SmMn2 O5
[87]), and the Ising chain magnet Ca3 Co2-x Mnx O6 [88]. Note that ferro-
electric polarization arising from such collinear spin ordering is generally

Fig. 3 (a) Schematic of a


cycloidal spin spiral, where
the spin (orange arrows)
rotates around an axis normal
to the propagation wave
vector Q. The induced
polarization P is normal to
both the spin rotation axis and
the Q. (b) Schematic of a
helical spin spiral where the
spin rotates around the wave
vector Q. The induced
polarization P is parallel to
the Q
602 J.-M. Hu and L.-Q. Chen

larger than that those from noncollinear spin ordering. For example, the
cycloidal spin spiral in o-TbMnO3 can be transformed into collinear spin
ordering by applying pressure, leading to an order of magnitude increase in
the spontaneous polarization [89].
(iii) The spin-dependent orbital hybridization between the 3d orbitals of mag-
netic transition metal (Co, Fe, Cr) and the 2p orbitals of ligand (O) via
spin-orbit coupling [90]. Examples include triangular-lattice antiferromag-
nets such as CuFeO2 [91] and CuCrO2 [92] and a staggered antiferromagnet
Ba2 CoGe2 O7 [93].

Ferroelectric antiferromagnets with lone pair-driven or geometrically driven


ferroelectricity are Type I multiferroics, while those with spin-driven ferroelectricity
belong to Type II. Table 3 summarizes the spontaneous/saturation polarization,
magnetization, and the onset temperatures for the ferroelectric order (TFE ) and
antiferromagnetic order (TAFM ) of some representative materials.

Ferroelectric Ferrimagnets
Ferroelectric ferrimagnets mainly include Fe3 O4 [103, 104], LuFe2 O4 [105], and
Pr1-x Cax MnO3 [106]. In these materials, ferroelectricity was thought to arise from
charge ordering ([107] is a brief review on this topic), which has been hotly debated.
For example, reports have suggested that the LuFe2 O4 , a prototypical example
of charge-ordering-driven ferroelectricity [105] with ferrimagnetic ordering, is
not ferroelectric [108–110]. However, in a (LuFeO3 )9 /(LuFe2 O4 )1 superlattice,
robust ferroelectricity has been demonstrated by out-of-plane piezoreponse force
microscopy at room temperature, while an evident magnetic hysteresis loop has
been directly recorded [111] at 250 K. Combining high-resolution transmission
electron microscopy imaging with first-principles calculations led to the suggestion
that the key ingredient for ferroelectricity is the severe rumpling imposed by the
neighboring LuFeO3 . It is such rumpling that drives the ferrimagnetic LuFe2 O4 into
a simultaneous ferroelectric state. Table 4 summarizes the spontaneous/saturation
polarization, magnetization, and the onset temperature for the ferroelectric order
(TFE ) and ferrimagnetism (TC ).

Electric Field Switching of Magnetization in Multiferroic BiFeO3 :


Status and Perspective

Electric field switching of magnetization in single-phase multiferroics is generally


weaker than it is in composite multiferroics/magnetoelectrics, partly because the
magnitude of magnetization in these single-phase materials (see Tables 2, 3,
and 4) is generally much smaller than that in ferromagnetic materials. However,
many of these single-phase multiferroics are excellent test beds for exploring new
phenomena and physics. In this section, we will briefly discuss the electric field
switching of magnetization in BiFeO3 , which is perhaps the most intensively studied
12 Magnetoelectrics and Multiferroics 603

Table 3 Properties of the ferroelectric-antiferromagnet-type single-phase multiferroics


Type I
or II
multi- Origin of fer- Spontaneous
Material ferroics roelectricity polarization TFE TAFM Reference
Pb(Fe1/2 Type I 6 s lone-pair 11.5 μC cm−2 380 K 145–160 K [94, 95]
Nb1/2 )O3 @RT in (TN )
ceramics
BiFeO3 Type I 6 s lone-pair 100 μC cm−2 1098 K 643 K (TN ) [69, 96,
@RT in bulk 97]
single crytals,
55 μC cm−2
@RT in
epitaxial thin
films
YMnO3 Type I Geometrically 5.5 μC cm−2 570– 70–130 K [98, 99]
driven @RT in bulk 990 K (TN )
single crystals
Hexagonal Type I Geometrically 5 μC cm−2 1050 K TN ∼ 400 K [71,
LuFeO3 driven @RT in bulk in film, 100,
single crytalsa TWFM ∼ 100 Kb 101]
BaNiF4 Type I Geometrically 6.7 μC cm−2 1320 K 50 K (TN ) [72, 75,
driven @RT in bulk (melts 102]
single crystals first)
Cr2 BeO4 Type II Spin-driven 3 × 10−4 μC 28 K 28 K (TN ) [81]
(noncollinear cm−2 @7 K @
spin ordering) 0.6 MV/md in
ceramics
TbMnO3 Type II Spin-driven 0.08 μC cm−2 27 K TN ∼ 41 K, [29]
(cycloidal @10 K in bulk TCYCL ∼ 27 Kc
spin spiral) single crystals
CaMn7 O12 Type II Spin-driven 0.287 μC 90 K 90 K (TN ) [82]
(noncollinear cm−2 @15 K
spin ordering) in bulk single
crystals,e 0.024
μC cm−2
@10 K in
ceramics
TbMn2 O5 Type II Spin-driven 0.04 μC cm−2 40 K 40 K (TN ) [30]
(collinear spin @3 K in bulk
ordering) single crystals
(continued)
604 J.-M. Hu and L.-Q. Chen

Table 3 (continued)
Type I
or II
multi- Origin of fer- Spontaneous
Material ferroics roelectricity polarization TFE TAFM Reference
TbMnO3 Type II Spin-driven 1 μC cm−2 20 K 20 K [89]
under (collinear spin @5 K @ (TE-AFM )f
high ordering) 5.2 GPa in bulk
pressure single crystals
(> 5 GPa)
CuCrO2 Type II Spin-driven 0.03 μC cm−2 24 K 24 K [92]
(p-d orbital @5 K @ 30
hybridization) MV/md in bulk
single crystals
a Estimated from structural analyses, not through direct measurement
b Will experience a second phase transition to weak ferromagnetism via spin canting below 100 K,
the TWFM
c Below T
CYCL , the sinusoidal AFM ordering (emerging when TCYCL < T < TN ) will transform into
a cycloidal spin spiral ordering. The latter induces ferroelectricity
d Thus it does not represent spontaneous polarization, unlike others
e It is still under debate whether CaMn O indeed exhibits such a surprisingly large spontaneous
7 12
polarization
f Under high pressure (>5 GPa), the sinusoidal AFM ordering will transform into a collinear E-type

AFM ordering, instead of transforming into the cycloidal spiral mentioned above (see temperature-
pressure phase diagram in Ref. [89])

Table 4 Properties of the ferroelectric-antiferromagnet-type single-phase multiferroics


Type I Origin
or II of ferro- Spontaneous Spontaneous
multi- electric- polariza- magneti-
Material ferroics ity tion zation TFE TC Reference
Fe3 O4 Type I Charge 5.0 μC 4.0 μB per 119–125 Ka 858 K [103,
order cm−2 @ Fe atom 112]
4.2 K @ 125 K
(LuFeO3 )9 / Type I Charge 6.0 μC 2.0 μB per >700 K 281 K [111]
(LuFe2 O4 )1 b order cm−2 @ Fe atom
RT @ 50 K
a Ferroelectricityis thought to emerge below the Verwey transition temperature TV , below which
the manganite transforms from a metal to insulator
b The superlattices were grown by reactive-oxide molecular-beam epitaxy on (111)

(ZrO2 )0.905 (Y2 O3 )0.095 (or 9.5 mol% yttria-stabilized zirconia, YSZ) substrates

multiferroic material. Since a famous 2003 Science paper [96] that reports a large
spontaneous polarization and an arguably large [113] spontaneous magnetization
in BiFeO3 thin films, research efforts into BiFeO3 have remained the center stage
for multiferroics and magnetoelectrics. In fact, there exist no less than seven (2006–
2018) comprehensive review articles that specifically discuss BiFeO3 [69, 114–120]
including some very comprehensive ones [118–120]. Here we provide a briefing of
12 Magnetoelectrics and Multiferroics 605

some of the seminal research efforts on the electric field control of magnetization in
BiFeO3 .

BiFeO3 Bulk Single Crystals


BiFeO3 has a long-range antiferromagnetic spin cycloid with a period of about
64 nm [121, 122] in rhombohedral bulk crystals (Fig. 4a). Canted antiferromagnetic
spins give rise a weak magnetic moment Mc via the Dzyaloshinskii-Moriya
interaction [125, 126], demonstrated through both first-principles calculations [124]
and experiments [127]. The local magnetic moment Mc , however, averages out over
one cycloid period; thus the BiFeO3 bulk crystal does not exhibit a macroscopic
magnetization. Remarkably, in high-quality single-domain BiFeO3 crystals, it has
been experimentally demonstrated [123] that the cycloidal plane (within which the
spin rotates, also defined as the magnetic easy plane) contains the spontaneous
polarization along one of the 111 directions, the local net magnetic moment Mc ,


and the cycloidal propagation vector Q along one of the 110 directions. Notably,
as the polarization is rotated by 71◦ through the application of an electric field, the

Fig. 4 (a) Schematic of an antiferromagnetic spin cycloid. The canted antiferromagnetic spins
(blue and green, representing two sublattices) rotate within the plane defined by the polarization
P and the cycloidal propagation vector Q. They also induce a net magnetic moment Mc (orange)
which is also within that plane and averaged out to zero over one cycloid period of λ = 64 nm.
(b) Schematic of the magnetic easy plane (antiferromagnetic plane) containing the rotating
spins (double-headed arrows), the P, and the Q in BiFeO3 bulk crystals. (c) Schematic of G-
type antiferromagnetism (left) and the 3D view of the Mc induced by canted antiferromagnetic
spins via Dzyaloshinskii-Moriya interaction. (d) Relationship between the P, the Mc , and the
antiferromagnetic vector L (parallel to the magnetic easy plane) in BiFeO3 epitaxial thin films in
which the spin cycloid is destroyed by epitaxial strains or/and reduced thickness. (a, b), Redrawn
based on Ref. [123]. (c) Redrawn based on Ref. [124]
606 J.-M. Hu and L.-Q. Chen

magnetic easy plane also rotates (Fig. 4b), as demonstrated experimentally by two
different research groups [128, 129]. If decorating the BiFeO3 single crystal with a
soft magnetic thin film (such as Permalloy), the electric field-induced reorientation
of magnetic easy plane can further switch the magnetic anisotropy of the soft
magnetic thin film [128].

Epitaxial BiFeO3 Thin Films


The long-range spin cycloid in BiFeO3 can be transformed into homogenous antifer-
romagnetic ordering by applying a strong (>18 T) magnetic field [28, 130–132]. In
epitaxial BiFeO3 thin films, it has long been suggested [133] that the epitaxial strain
may destroy the spin cycloid and thus unleash the hidden weak ferromagnetism.
Combined experimental and theoretical analyses [134] suggested that large tensile
and compressive strain can transform the spin cycloid into homogenous pseudo-
collinear antiferromagnetic ordering and that a moderate tensile strain can stabilize
a new type of spin cycloid with a propagation vector along one of the 110 directions.
A relatively complete temperature-strain phase diagram for BiFeO3 can be found in
[118].
Furthermore, in BiFeO3 thin film with pseudo-collinear G-type antiferromag-
netic ordering (slightly canted spins, see Fig. 4c), where spin cycloid is destroyed
by strain or possibly by the reduced thickness [69], first-principles calculations have
predicted that the spontaneous polarization is perpendicular to the magnetic easy
plane, which contains the canted antiferromagnetic moments MFe1 and MFe2 as well
as their corresponding antiferromagnetic vector L = MFe1 -MFe2 and canted weak
magnetization Mc = MFe1 + MFe2 (Mc and L are mutually orthogonal), as shown
in Fig. 4d. In this case, as polarization is switched by 71◦ or 109◦ , both the magnetic
easy plane and the Mc will be switched. Remarkably, the coupling between polar-
ization and the magnetic easy plane has been experimentally demonstrated [135] by
simultaneously visualizing the ferroelectric and antiferromagnetic domains using
piezoresponse force microscopy and X-ray photoemission electron microscopy
(X-PEEM), respectively. In contrast, a direct observation of polarization switching-
induced reorientation of Mc is still lacking. Nevertheless, the Mc reorientation has
been indirectly demonstrated by the observation of local magnetization switching
in a 2.5-nm-thick polycrystalline Co0.9 Fe0.1 film atop BiFeO3 [136–138], where the
local magnetization in the Co0.9 Fe0.1 film and the Mc are locked through atomistic
Heisenberg-type exchange coupling [139]. Notably, a full 180◦ reversal of local
magnetization in the Co0.9 Fe0.1 film has been directly observed by X-PEEM [138]
after the application and removal of an electric field across the Co0.9 Fe0.1 /BiFeO3
heterostructure. This was thought to be induced by a two-step 180◦ reversal of the
Mc accompanying a unique two-step polarization switching at the BiFeO3 surface.
The ability to electrically switch the magnetization in a magnetic film atop BiFeO3
offers interesting application as novel exchange coupling-mediated electric-write
magnetic memories; see a perspective article [140] and a review on magnetoelectric
devices for spintronics [141].
12 Magnetoelectrics and Multiferroics 607

A Brief Future Perspective

Electric Field Control of Spin Cycloid in BiFeO3 Thin Films: From the Control
of Collinear Magnetism to Noncollinear Magnetism
Electric field control of spin cycloid (and magnetic easy plane) has been experi-
mentally demonstrated in BiFeO3 bulk crystals [123] but not yet in BiFeO3 thin
films, although epitaxial strain-induced switching of spin cycloid in BiFeO3 thin
films has been experimentally demonstrated [134]. The coupling among the oxygen
octahedral tilts, polarization, and the spin cycloid, which has been discussed with
phenomenological approaches [142, 143], awaits experimental investigation.

Towards the Control of Magnetism in BiFeO3 Thin Films at the THz


Frequency: New Opportunities with Ultrafast Stimuli
Existing experimental reports on electric field control of magnetism (including spin
cycloid [123, 128, 129], canted magnetization Mc [136–139], and spin waves [144])
in BiFeO3 all use a static or low-frequency electric field. It remains unclear how
the magnetism or antiferromagnetism in BiFeO3 will respond to an ultrafast THz
electric field, which can be created by applying a femtosecond laser pulse to the
BiFeO3 via the photostriction effect [145]. Note that applying femtosecond laser
excitation to ferroelectrics has already led to the discovery of several new phenom-
ena that cannot be accessed through static/low-frequency electric fields, including,
for example, the polarization oscillation [146, 147] and up to 0.5% transitional
strains [148, 149] that are orders of magnitude larger than those obtained through
linear inverse piezoelectric effect. Furthermore, in a heterostructure consisting of
thin Ni film deposited onto BiFeO3 bulk substrates, it has been experimentally
demonstrated [150] that an optically induced strain at the BiFeO3 surface can further
modify the magnetic anisotropy of the Ni film, offering new opportunities as a
strain-mediated optical control of magnetism.

Composite Multiferroics and Magnetoelectrics

Terminology and Exiting Reviews

Composite multiferroics (also “multiferroic heterostructures” or “artificial multifer-


roics”) integrate magnetic and ferroelectric materials to produce magnetoelectric
effects that are absent in either the magnetic or ferroelectric phase. Compos-
ite magnetoelectrics, strictly speaking, represent composites integrating magnetic
and piezoelectric materials, enabling both the inverse (electric field control of
magnetization) and direct (magnetic field control of electric polarization) magne-
toelectric effects. However, composites integrating magnetic and plain-dielectric
(such as MgO, GdOx ) or single-phase magnetoelectric (such as Cr2 O3 ) materials
(see Fig. 5a), in which only the inverse magnetoelectric effect has so far been
608 J.-M. Hu and L.-Q. Chen

Fig. 5 (a) Classifying composite magnetoelectrics based on the functionality of the constituent
dielectric material, along with representative dielectric materials. The solid ellipses suggest the
corresponding mechanisms of inverse magnetoelectric effects. EC, exchange coupling; FE-AF,
ferroelectric antiferromagnet; ME-AF, magnetoelectric antiferromagnet. Redrawn based on Ref.
[151] http://creativecommons.org/licenses/by/4.0/. (b) Relationship between composite multifer-
roics and composite magnetoelectrics by a loose definition, see discussion in the main text

reported, can also be called as composite magnetoelectrics (also “magnetoelectric


heterostructures”) by a loose definition. Thus, we may conclude that composite
multiferroics represent a subset of composite magnetoelectrics (Fig. 5b).
In contrast to single-phase magnetoelectric and multiferroic materials, magne-
toelectric effects in these composite materials have no upper bound set by the
geometric mean of the electric permittivity and magnetic permeability. Rather,
the magnitudes of the magnetoelectric effects can be tailored by varying the
choices of materials and designing the geometry and microstructure of each
constituent materials and can reach the level of 10−5 s/m that is five orders of
magnitude larger than the record in single-phase magnetoelectrics and multiferroics
(i.e., ∼10−10 s/m in magnetoelectric TbPO4 [26, 152]). Furthermore, the magneto-
electric effects typically can emerge at room temperature. Overall, the design flex-
ibility and the ability to achieve large magnetoelectric effects at room temperature
make composite multiferroics/magnetoelectrics more attractive for potential device
applications.
There exist a series of comprehensive review articles on composite multiferroics
[151, 153–159], brief perspectives [160–165], a few topical reviews on electric field
control of magnetism [117, 166–170], and the magnetoelectric device applications
[34, 141, 171–174]. Here, we will briefly discuss the mechanisms of magnetoelec-
tric effects and the experimental data on magnetoelectric coefficients in composite
magnetoelectrics as well as their potential device applications. For a more detailed
discussion of the field, see Refs. [151, 159].
12 Magnetoelectrics and Multiferroics 609

Mechanisms and Application of Magnetoelectric Effects and


Experimental Data of Magnetoelectric Coefficients in Composite
Magnetoelectrics

Magnetoelectric effects enable a mutual conversion of magnetic and electric energy


in the absence of an electric current, while the magnetoelectric coefficient represents
a measure on the degree of energy conversion. In general, magnetoelectric effects in
composite magnetoelectrics arise from the coupling among the four fundamental
degrees of freedom (spin, charge, lattice, orbit) across the magnetic/ferroelec-
tric (dielectric) interface through the exchange of magnetic, electric, and elastic
energy or/and mass. From a phenomenological perspective, different mechanisms
of magnetoelectric effects indicate different energy conversion pathways. For
example, a strain-mediated magnetic field control of electric polarization (direct
magnetoelectric effect) in magnetic/piezoelectric composites is achieved through a
conversion of magnetic energy to elastic energy (via magnetostrictive effect in the
magnetic phase) then to electric energy (via piezoelectric effect in the piezoelectric
phase) and vice versa for the corresponding inverse magnetoelectric effect.

Direct Magnetoelectric Effect


The interfacial strain transfer mechanism mentioned above remains the only
reported mechanism for a direct magnetoelectric effect in composite magneto-
electrics. Direct magnetoelectric effects are attractive mainly for applications in
the form of low-cost and energy-efficient magnetic field sensors [171], also known
as “magnetoelectric sensors” in the literature. Of particular interest is the use of
magnetoelectric sensors, instead of the currently used superconducting quantum
interference device (SQUID) sensors which are cumbersome and expensive, to
detect weak (10 f T – 1 pT) and low-frequency (10−2 – 103 Hz) magnetic fields
generated from the electrical activities of human organs [174]. This could lead
to the development of a series of household devices for biomedical diagnosis,
yet it remains a challenging goal because appropriate composite materials are
still lacking. Specifically, Fig. 6a, b shows a relatively extensive experimental
dataset of direct magnetoelectric voltage coefficient (α HV = E/(μ0 HAC )) that we
compiled [151] for bulk and thin-film composite magnetoelectrics of different phase
connectivity [175] (see schematics in Fig. 6c). The data of the α HV magnitudes are
presented as a function of the frequency (fAC ) of the driving AC magnetic field
μ0 HAC (Fig. 6a) and the magnitude of the bias magnetic field μ0 HDC (Fig. 6b).
A high α HV under low fAC and small μ0 HDC is desirable for magnetoelectric
sensor applications [151, 176]. However, as shown in Fig. 6a, there are no available
experimental data of α HV in the frequency range 0 < fAC < 1 Hz.

Inverse Magnetoelectric Effect


The inverse magnetoelectric effect generally refers to the modulation of magnetism
with an electric field, with potential device applications to energy-efficient electric-
610 J.-M. Hu and L.-Q. Chen

Fig. 6 Complied experimental datasets showing the direct magnetoelectric voltage coefficientα HV
as a function of (a), the frequency of the driving AC magnetic field, and (b), the magnitude of
the bias magnetic field, both of which are necessary for the experimental measurement of α HV .
Available data in both bulk and thin-film composites of different phase connectivity are compiled.
Adapted from Ref. [151] http://creativecommons.org/licenses/by/4.0/. PC, particulate composites;
HC, horizontal composites; VC, vertical composites (see schematics in (c))

write magnetic-read memories [140, 177–179] and logic [180–184], electric field
tunable radiofrequency or microwave devices [158], etc. For example, electric field-
driven magnetization switching in the free layer of a magnetic tunnel junction (MTJ)
allows us to electrically toggle the electrical resistance of the entire junction between
high and low value, representing the bit information “1” and “0,” respectively.
The inverse magnetoelectric effect in composites magnetoelectrics may occur
through different mechanisms, depending on the functionality of the constituent
dielectric materials. Below we briefly discuss these mechanisms in composite
magnetoelectrics with a “plain” dielectric, piezoelectric, ferroelectric, ferroelectric
antiferromagnet, or a magnetoelectric antiferromagnet, respectively. We use a
layered magnetoelectric (Fig. 7) for discussion.
Plain Dielectric (Oxygen Ion Insulator). In the case of a “plain” dielectric
with low concentration of ion-conducting defects (such as MgO and HfO2 , see
Fig. 5a for representative materials), the inverse magnetoelectric effect typically
occurs through the modulation of spin-polarized charge densities (ns ≈ nη)
at the magnetic/dielectric interface as voltage is applied. Here n = ε0 εr E is
the change of the screened charges densities with ε0 and εr denoting the vacuum
and relative permittivity, respectively. The electric field-induced change of interface
12 Magnetoelectrics and Multiferroics 611

Fig. 7 Schematics of inverse magnetoelectric effects in magnetoelectric heterostructures through


electrically modulated (a) spin-polarized charge densities, (b) degree of interfacial oxidation via
oxygen vacancy accumulation/depletion, (c) strain, and (d) interfacial exchange coupling. In (a),
an enhanced (reduced) surface electron densities of the dielectric layer will reduce (enhance) the
interfacial electron densities of the magnetic layer through electrostatic interaction, further shifting
the Fermi level at the interface region of the magnet to a lower (higher) level, as shown in the top
(bottom) panel. In (b), the MOx layer could affect the perpendicular magnetic anisotropy through
the modulation of interfacial M-O orbital hybridization or/and imposing an interfacial exchange
bias field, the details of which await further clarification. In (c), the electric field-induced strains are
transferred to the overlaid magnet and modulate magnetization through magnetoelastic coupling.
Different from the other mechanisms discussed here, strain-magnetization coupling is long-range.
In (d), only one single antiferromagnetic domain with perpendicular sublattice magnetization (see
arrows in the bottom layer) is shown for simplicity, which is also made largely based on a Cr2 O3 -
based magnetoelectric heterostructure. Details of inverse magnetoelectric effects in BiFeO3 -based
or YMnO3 -based heterostructures may vary. (a–d) Reprinted with permission from Ref. [151]
http://creativecommons.org/licenses/by/4.0/

charge densities will shift the Fermi level at the interface, further changing the
interfacial spin polarization P = (D↑ − D↓ )/(D↑ + D↓ ), where D↑ (D↓ ) indicates
the densities of states for the spin-up (down) electrons, as schematically shown in
Fig. 7a. Notably, inverse magnetoelectric effect through such mechanism, which
we term as charge densities, can be directly observed [185–188] in a magnetic
tunnel junction such as CoFeB(pinned layer)/MgO/CoFeB (free layer) without
integrating additional dielectric materials. This is an important advantage for device
applications. However, such charge density-mediated inverse magnetoelectric effect
is generally very weak (see our compiled experimental dataset in Fig. 8a) and
usually causes an electric field-controlled interface magnetic anisotropy of less than
0.1 pJ/m2 per unit electric field (1 V/m), which is too small to switch a magnetization
with a sufficiently high thermal stability (>40 kB T) [189].
612 J.-M. Hu and L.-Q. Chen

Fig. 8 (a) Compiled experimental datasets showing the inverse magnetoelectric coefficient α E
(the slope) in magnetoelectric heterostructures, where the horizontal and vertical axes represent the
driving electric field and the electrically induced change of magnetization, respectively. The data
are classified by the corresponding mechanisms of inverse magnetoelectric effects as discussed
in Fig. 7, including charge densities (CD), interfacial oxidation (IO), exchange coupling (EC),
and strain transfer (ST). (b) Corresponding thickness of the constituent magnetic layer in a
chronological order. (a–b) Reprinted with permission from Ref. [151] http://creativecommons.org/
licenses/by/4.0/

Plain Dielectric (Oxygen Ion Conductor). If the dielectric layer is a good oxygen
ion conductor (such as GdOx ) and if the magnetic layer (M = Co, Fe, CoFe)
has a perpendicular magnetic anisotropy (PMA) that is sensitive to the degree of
interface oxidation (i.e., the x in an interfacial MOx layer), it is then possible to
tune the PMA by applying a voltage to tune the O2− concentration at the interface
and hence the degree of interface oxidation. The inverse magnetoelectric effect
through such mechanism, which we term as interfacial oxidation (see schematic
in Fig. 7b), has been observed mainly in Co/GdOx heterostructures [190–193],
and it can enable a giant electric field-controlled interface magnetic anisotropy of
more than 10 pJ V−1 m−1 , over 100 times larger than that in the case of charge
densities mechanism. This can be also seen from the approximately 100 times
larger inverse magnetoelectric coefficient as shown in our complied experimental
dataset (10−8 s/m vs. 10−10 s/m, see Fig. 8a). However, the speed of the inverse
magnetoelectric effect through such an interfacial oxidation mechanism may be low
compared to that through the charge density mechanism, because of intrinsically
slow oxygen ion migration. In addition, the back-and-forth oxygen-ion migration
between the upper and lower surfaces of the dielectric may accelerate dielectric
degradation. Thus, the potential device application of such interface oxidation
mechanism remains an open question.
Piezoelectric or Ferroelectric Dielectric. If the dielectric layer is piezoelectric ,
magnetism can be tuned, through magnetoelastic coupling, by strain imparted by
the piezoelectric layer as voltage is applied (see schematic in Fig. 7c), but it can
12 Magnetoelectrics and Multiferroics 613

also be tuned through the charge density mechanism mentioned previously. If the
piezoelectric layer is also ferroelectric, it is possible to obtain larger strains through
ferroelectric domain switching or a structural phase transition and larger electrically
induced change of interfacial charge densities due to the larger relative permittivity
εr in the ferroelectric. Both will lead to larger inverse magnetoelectric effects.
Furthermore, strain-mediated electric field-driven full magnetization reversal [194,
195] and unidirectional magnetic domain-wall motion [196] have recently been
computationally demonstrated at room temperature, providing exciting application
potential for electric-write magnetic-read memories. However, microfabrication of
some piezoelectric or ferroelectric materials may not be compatible with Si CMOS
(complementary metal-oxide-semiconductor) processing, and the fatigue problem
of ferroelectrics is also a potential issue.
Ferroelectric (or Magnetoelectric) Antiferromagnet. If the ferroelectric is also
antiferromagnetic such as BiFeO3 , the modulation of magnetism in the magnetic
layer can also be achieved, through Heisenberg-type exchange coupling, by elec-
trically tuning the antiferromagnetic order through reorientation of the magnetic
easy plane or/and the canted magnetization (see details in Sect. “Electric Field
Switching of Magnetization in Multiferroic BiFeO3 : Status and Perspective”).
Notably, exchange coupling-mediated electric field-driven 180◦ net magnetization
reversal has been experimentally demonstrated at room temperature [138]. However,
the high leakage current of BiFeO3 as well as its incompatibility with Si CMOS
processing in terms of microfabrication (e.g., specific oxide substrates must be
used for film deposition) undermines the application potential of BiFeO3 -based
magnetoelectric heterostructures. Such an exchange coupling mechanism has also
been observed in Cr2 O3 -based magnetoelectric heterostructures at room tempera-
ture [197]. Cr2 O3 is a magnetoelectric antiferromagnet, where the antiferromagnetic
domains as well as the uncompensated surface magnetic moment can be reversed by
applying an electric field (see schematic in Fig. 7d). Thus, the interfacial magnetic
moments in the overlaid magnet can be tuned through exchange coupling. Cr2 O3
has excellent dielectric properties with high dielectric breakdown field of 109 V/m
at room temperature, which is an advantage for potential device application,
although it remains a challenge to achieve such dielectric properties in Cr2 O3
thin films. Moreover, the electric field switching of antiferromagnetic domains in
Cr2 O3 requires simultaneous application of a static magnetic field to lift the time-
reversal symmetry. This will be a potential disadvantage for high-density device
applications, as it could complicate the device design by increasing the noise level
and possibly causing cross talk among neighboring device units.
Among the different mechanisms discussed above, only the strain-magnetization
coupling is long-range, while all the others are interface effects, consistent with our
complied experimental dataset showing the thickness of the constituent magnetic
layer in the case of different mechanisms (see Fig. 8b). Such long-range strain-
magnetization coupling partially leads to the fact that strain-mediated inverse
magnetoelectric effect is generally larger than the other types of effects, consistent
with the experimental data shown in Fig. 8a. For more detailed discussions on these
different mechanisms as well as remaining open questions, see He et al. [151].
614 J.-M. Hu and L.-Q. Chen

Acknowledgments This work was supported by a start-up fund from the University of Wisconsin-
Madison (J.-M.H.) and partially by the National Science Foundation under the grant no. DMR-
1744213 (Chen) and partially by the Army Research Office under the grant number W911NF-17-
1-0462 (J.-M.H. and L.-Q.C.). The authors acknowledge Xin Zou for helping draw some of the
schematics.

References
1. Curie, P.: Sur la symétrie dans les phénomènes physiques, symétrie d’un champ électrique et
d’un champ magnétique. J. Phys. Theor. Appl. 3(1), 393–415 (1894)
2. Debye, P.: Bemerkung zu einigen neuen Versuchen über einen magneto-elektrischen Richtef-
fekt. Z. Phys. 36(4), 300–301 (1926)
3. O’Dell, T.H.: The Electrodynamics of Magneto-Electric Media. North-Holland Publishing
Company, Amsterdam (1970)
4. Landau, L.D., Lifshitz, E.M.: Course of Theoretical Physics: Electrodynamics of Continuous
Media. Pergamon Press, Oxford (1957).(English Transl. in 1960)
5. Röntgen, W.C.: Ueber die durch Bewegung eines im homogenen electrischen Felde befind-
lichen Dielectricums hervorgerufene electrodynamische Kraft. Ann. Phys. 271(10), 264–270
(1888)
6. Wilson, H.A.: On the electric effect of rotating a dielectric in a magnetic field. Philos. Trans.
R. Soc. Lond. A. 204, 121–137 (1905)
7. Hou, S.L., Bloembergen, N.: Paramagnetoelectric effects in NiSO4 ·H2 O. Phys. Rev. 138(4A),
A1218–A1226 (1965)
8. Dzyaloshinskii, I.: On the magneto-electrical effect in antiferromagnets. Sov. Phys. JETP.
10(3), 628–629 (1960)
9. Astrov, D.: The magnetoelectric effect in antiferromagnetics. Sov. Phys. JETP. 11(3), 708–
709 (1960)
10. Folen, V.J., Rado, G.T., Stalder, E.W.: Anisotropy of the magnetoelectric effect in Cr2 O3 .
Phys. Rev. Lett. 6(11), 607–608 (1961)
11. Rado, G.T., Folen, V.J.: Observation of the magnetically induced magnetoelectric effect and
evidence for antiferromagnetic domains. Phys. Rev. Lett. 7(8), 310–311 (1961)
12. Shtrikman, S., Treves, D.: Observation of the magnetoelectric effect in Cr2 O3 powders. Phys.
Rev. 130(3), 986–988 (1963)
13. Al’Shin, B., Astrov, D.: Magnetoelectric effect in titanium oxide Ti2 O3 . Sov. Phys. JETP.
17(4), 809–811 (1963)
14. Rado, G.T.: Observation and possible mechanisms of magnetoelectric effects in a ferromag-
net. Phys. Rev. Lett. 13(10), 335–337 (1964)
15. Ascher, E., Rieder, H., Schmid, H., Stössel, H.: Some properties of ferromagnetoelectric
nickel-iodine boracite, Ni3 B7 O13 I. J. Appl. Phys. 37(3), 1404–1405 (1966)
16. Mercier, M., Gareyte, J., Bertaut, E.: Une nouvelle famille de corps magnetoelectriques-
LiMPO4 (M= Mn, Co, Ni). Comptes Rendus. 264(13), 979 (1967)
17. Rivera, J.-P.: The linear magnetoelectric effect in LiCoPO4 revisited. Ferroelectrics. 161(1),
147–164 (1994)
18. Fischer, E., Gorodetsky, G., Hornreich, R.M.: A new family of magnetoelectric materials:
A2 M4 O9 (A=Ta, Nb; M=Mn, Co). Solid State Commun. 10(12), 1127–1132 (1972)
19. Watanabe, T., Kohn, K.: Magnetoelectric effect and low temperature transition of PbFe0.
5Nb0. 5O3 single crystal. Phase Transit. Multinatl. J. 15(1), 57–68 (1989)
20. Schmid, H.: On a magnetoelectric classification of materials. Int. J. Magnetism. 4(4), 337–361
(1973)
21. Fiebig, M.: Revival of the magnetoelectric effect. J. Phys. D. Appl. Phys. 38(8), R123–R152
(2005)
12 Magnetoelectrics and Multiferroics 615

22. Brown, W.F.: Micromagnetics: Interscience Tracts on Physics and Astronomy, 18. J. Wiley,
New York, London (1963)
23. Brown, W.F., Hornreich, R.M., Shtrikman, S.: Upper bound on the magnetoelectric suscepti-
bility. Phys. Rev. 168(2), 574–577 (1968)
24. Schmid, H.: Multi-ferroic magnetoelectrics. Ferroelectrics. 162(1–4), 317–338 (1994)
25. Borisov, P., Ashida, T., Nozaki, T., Sahashi, M., Lederman, D.: Magnetoelectric properties of
500-nm Cr2 O3 films. Phys. Rev. B. 93(17), 174415 (2016)
26. Rado, G.T., Ferrari, J.M., Maisch, W.G.: Magnetoelectric susceptibility and magnetic sym-
metry of magnetoelectrically annealed TbPO4 . Phys. Rev. B. 29(7), 4041–4048 (1984)
27. Chun, S.H., Chai, Y.S., Jeon, B.-G., Kim, H.J., Oh, Y.S., Kim, I., Kim, H., Jeon, B.J., Haam,
S.Y., Park, J.-Y., Lee, S.H., Chung, J.-H., Park, J.-H., Kim, K.H.: Electric field control of
nonvolatile four-state magnetization at room temperature. Phys. Rev. Lett. 108(17), 177201
(2012)
28. Tokunaga, M., Akaki, M., Ito, T., Miyahara, S., Miyake, A., Kuwahara, H., Furukawa,
N.: Magnetic control of transverse electric polarization in BiFeO3 . Nat. Commun. 6, 5878
(2015)
29. Kimura, T., Goto, T., Shintani, H., Ishizaka, K., Arima, T., Tokura, Y.: Magnetic control of
ferroelectric polarization. Nature. 426(6962), 55–58 (2003)
30. Hur, N., Park, S., Sharma, P.A., Ahn, J.S., Guha, S., Cheong, S.W.: Electric polarization
reversal and memory in a multiferroic material induced by magnetic fields. Nature. 429(6990),
392–395 (2004)
31. Eerenstein, W., Mathur, N.D., Scott, J.F.: Multiferroic and magnetoelectric materials. Nature.
442(7104), 759–765 (2006)
32. Dong, S., Liu, J.-M., Cheong, S.-W., Ren, Z.: Multiferroic materials and magnetoelectric
physics: symmetry, entanglement, excitation, and topology. Adv. Phys. 64(5–6), 519–626
(2015)
33. Fiebig, M., Lottermoser, T., Meier, D., Trassin, M.: The evolution of multiferroics. Nat. Rev.
Mater. 1, 16046 (2016)
34. Scott, J.F.: Applications of magnetoelectrics. J. Mater. Chem. 22(11), 4567–4574 (2012)
35. Aizu, K.: Possible species of ferromagnetic, ferroelectric, and ferroelastic crystals. Phys. Rev.
B. 2(3), 754–772 (1970)
36. Aizu, K.: Possible species of “ferroelastic” crystals and of simultaneously ferroelectric and
ferroelastic crystals. J. Phys. Soc. Jpn. 27(2), 387–396 (1969)
37. Newnham, R.E.: Domains in minerals. Am. Miner. 59(9–10), 906–918 (1974)
38. Cheong, S.W., Mostovoy, M.: Multiferroics: a magnetic twist for ferroelectricity. Nat. Mater.
6(1), 13–20 (2007)
39. Ramesh, R., Spaldin, N.A.: Multiferroics: progress and prospects in thin films. Nat. Mater.
6(1), 21–29 (2007)
40. Tokura, Y.: Multiferroics – toward strong coupling between magnetization and polarization
in a solid. J. Magn. Magn. Mater. 310(2), 1145–1150 (2007)
41. Martin, L., Crane, S.P., Chu, Y.H., Holcomb, M.B., Gajek, M., Huijben, M., Yang, C.H.,
Balke, N., Ramesh, R.: Multiferroics and magnetoelectrics: thin films and nanostructures. J.
Phys. Condens. Matter. 20(43), 434220 (2008)
42. Spaldin, N.A., Cheong, S.-W., Ramesh, R.: Multiferroics: past, present, and future. Phys.
Today. 63(10), 38–43 (2010)
43. Lawes, G., Srinivasan, G.: Introduction to magnetoelectric coupling and multiferroic films. J.
Phys. D. Appl. Phys. 44(24), 243001 (2011)
44. Spaldin, N.A., Ramesh, R.: Advances in magnetoelectric multiferroics. Nat. Mater. 18(3),
203–212 (2019)
45. Khomskii, D.: Classifying multiferroics: mechanisms and effects. Physics. 2(20) (2009)
46. Kimura, T., Lawes, G., Ramirez, A.P.: Electric polarization rotation in a hexaferrite with long-
wavelength magnetic structures. Phys. Rev. Lett. 94(13), 137201 (2005)
47. Ishiwata, S., Taguchi, Y., Murakawa, H., Onose, Y., Tokura, Y.: Low-magnetic-field control
of electric polarization vector in a helimagnet. Science. 319(5870), 1643–1646 (2008)
616 J.-M. Hu and L.-Q. Chen

48. Chun, S.H., Chai, Y.S., Oh, Y.S., Jaiswal-Nagar, D., Haam, S.Y., Kim, I., Lee, B., Nam, D.H.,
Ko, K.-T., Park, J.-H., Chung, J.-H., Kim, K.H.: Realization of giant magnetoelectricity in
helimagnets. Phys. Rev. Lett. 104(3), 037204 (2010)
49. Kitagawa, Y., Hiraoka, Y., Honda, T., Ishikura, T., Nakamura, H., Kimura, T.:
Low-field magnetoelectric effect at room temperature. Nat. Mater. 9(10), 797–802
(2010)
50. Wang, F., Zou, T., Yan, L.-Q., Liu, Y., Sun, Y.: Low magnetic field reversal of electric
polarization in a Y-type hexaferrite. Appl. Phys. Lett. 100(12), 122901 (2012)
51. Okumura, K., Haruki, K., Ishikura, T., Hirose, S., Kimura, T.: Multilevel magnetization
switching by electric field in c-axis oriented polycrystalline Z-type hexaferrite. Appl. Phys.
Lett. 103(3), 032906 (2013)
52. Chai, Y.S., Kwon, S., Chun, S.H., Kim, I., Jeon, B.-G., Kim, K.H., Lee, S.: Electrical control
of large magnetization reversal in a helimagnet. Nat. Commun. 5, 4208 (2014)
53. Hirose, S., Haruki, K., Ando, A., Kimura, T.: Mutual control of magnetization and
electrical polarization by electric and magnetic fields at room temperature in Y-type BaS-
rCo2−x Znx Fe11 AlO22 ceramics. Appl. Phys. Lett. 104(2), 022907 (2014)
54. Shen, S., Chai, Y., Sun, Y.: Nonvolatile electric-field control of magnetization in a Y-type
hexaferrite. Sci. Rep. 5, 8254 (2015)
55. Kimura, T.: Magnetoelectric hexaferrites. Annu. Rev. Condens. Matter Phys. 3(1), 93–110
(2012)
56. Hill, N.A.: Why are there so few magnetic ferroelectrics? J. Phys. Chem. B. 104(29), 6694–
6709 (2000)
57. Sugawara, F., Iiida, S., Syono, Y., Akimoto, S.-I.: Magnetic properties and crystal distortions
of BiMnO3 and BiCrO3 . J. Phys. Soc. Jpn. 25(6), 1553–1558 (1968)
58. Moreira dos Santos, A., Parashar, S., Raju, A.R., Zhao, Y.S., Cheetham, A.K., Rao, C.N.R.:
Evidence for the likely occurrence of magnetoferroelectricity in the simple perovskite,
BiMnO3 . Solid State Commun. 122(1), 49–52 (2002)
59. Kimura, T., Kawamoto, S., Yamada, I., Azuma, M., Takano, M., Tokura, Y.: Magnetocapaci-
tance effect in multiferroic BiMnO3 . Phys. Rev. B. 67(18), 180401 (2003)
60. Eerenstein, W., Morrison, F.D., Scott, J.F., Mathur, N.D.: Growth of highly resistive BiMnO3
films. Appl. Phys. Lett. 87(10), 101906 (2005)
61. Gajek, M., Bibes, M., Fusil, S., Bouzehouane, K., Fontcuberta, J., Barthelemy, A.E., Fert, A.:
Tunnel junctions with multiferroic barriers. Nat. Mater. 6(4), 296–302 (2007)
62. Shvartsman, V.V., Borisov, P., Kleemann, W., Kamba, S., Katsufuji, T.: Large off-diagonal
magnetoelectric coupling in the quantum paraelectric antiferromagnet EuTiO3 . Phys. Rev. B.
81(6), 064426 (2010)
63. Lee, J.H., Fang, L., Vlahos, E., Ke, X., Jung, Y.W., Kourkoutis, L.F., Kim, J.-W., Ryan, P.J.,
Heeg, T., Roeckerath, M., Goian, V., Bernhagen, M., Uecker, R., Hammel, P.C., Rabe, K.M.,
Kamba, S., Schubert, J., Freeland, J.W., Muller, D.A., Fennie, C.J., Schiffer, P., Gopalan, V.,
Johnston-Halperin, E., Schlom, D.G.: A strong ferroelectric ferromagnet created by means of
spin-lattice coupling. Nature. 466(7309), 954–958 (2010)
64. Fennie, C.J., Rabe, K.M.: Magnetic and electric phase control in epitaxial EuTiO3 from first
principles. Phys. Rev. Lett. 97(26), 267602 (2006)
65. von Wartburg, W.: The magnetic structure of magnetoelectric nickel–iodine boracite
Ni3 B7 O13 I. Phys. Status Solidi A. 21(2), 557–568 (1974)
66. Chiba, H., Atou, T., Syono, Y.: Magnetic and electrical properties of Bi1−x Srx MnO3 : hole-
doping effect on ferromagnetic perovskite BiMnO3 . J. Solid State Chem. 132(1), 139–143
(1997)
67. Chi, Z.H., Xiao, C.J., Feng, S.M., Li, F.Y., Jin, C.Q., Wang, X.H., Chen, R.Z., Li, L.T.:
Manifestation of ferroelectromagnetism in multiferroic BiMnO3 . J. Appl. Phys. 98(10),
103519 (2005)
68. Smolenskii, G., Agranovskaya, A., Popov, S., Isupov, V.: New ferroelectrics of complex
composition. 2. Pb2 Fe3+ NbO6 and Pb2 YbNbO6 . Soviet physics Tech. Phys. 3, 1981–1982
(1958)
12 Magnetoelectrics and Multiferroics 617

69. Catalan, G., Scott, J.F.: Physics and applications of bismuth ferrite. Adv. Mater. 21(24), 2463–
2485 (2009)
70. Fiebig, M., Lottermoser, T., Frohlich, D., Goltsev, A.V., Pisarev, R.V.: Observation of coupled
magnetic and electric domains. Nature. 419(6909), 818–820 (2002)
71. Wang, W., Zhao, J., Wang, W., Gai, Z., Balke, N., Chi, M., Lee, H.N., Tian, W., Zhu, L.,
Cheng, X., Keavney, D.J., Yi, J., Ward, T.Z., Snijders, P.C., Christen, H.M., Wu, W., Shen, J.,
Xu, X.: Room-temperature multiferroic hexagonal LuFeO3 films. Phys. Rev. Lett. 110(23),
237601 (2013)
72. Eibschütz, M., Guggenheim, H.J., Wemple, S.H., Camlibel, I., DiDomenico, M.: Ferroelec-
tricity in BaM2+ F4 . Phys. Lett. A. 29(7), 409–410 (1969)
73. Fox, D.L., Scott, J.F.: Ferroelectrically induced ferromagnetism. J. Phys. C: Solid State Phys.
10(11), L329 (1977)
74. Fox, D.L., Tilley, D.R., Scott, J.F., Guggenheim, H.J.: Magnetoelectric phenomena in
BaMnF4 and BaMn0.99 Co0.01 F4 . Phys. Rev. B. 21(7), 2926–2936 (1980)
75. Scott, J.F., Blinc, R.: Multiferroic magnetoelectric fluorides: why are there so many magnetic
ferroelectrics? J. Phys. Condens. Matter. 23(11) (2011)
76. Yoshinori, T., Shinichiro, S., Naoto, N.: Multiferroics of spin origin. Rep. Prog. Phys. 77(7),
076501 (2014)
77. Tokura, Y., Seki, S.: Multiferroics with spiral spin orders. Adv. Mater. 22(14), 1554–1565
(2010)
78. Kimura, T.: Spiral magnets as magnetoelectrics. Annu. Rev. Mater. Res. 37(1), 387–413
(2007)
79. Katsura, H., Nagaosa, N., Balatsky, A.V.: Spin current and magnetoelectric effect in non-
collinear magnets. Phys. Rev. Lett. 95(5), 057205 (2005)
80. Mostovoy, M.: Ferroelectricity in spiral magnets. Phys. Rev. Lett. 96(6), 067601 (2006)
81. Newnham, R.E., Kramer, J.J., Schulze, W.A., Cross, L.E.: Magnetoferroelectricity in
Cr2 BeO4 . J. Appl. Phys. 49(12), 6088–6091 (1978)
82. Johnson, R.D., Chapon, L.C., Khalyavin, D.D., Manuel, P., Radaelli, P.G., Martin, C.: Giant
improper ferroelectricity in the ferroaxial magnet CaMn7 O12 . Phys. Rev. Lett. 108(6), 067201
(2012)
83. Terada, N., Glazkova, Y.S., Belik, A.A.: Differentiation between ferroelectricity and thermally
stimulated current in pyrocurrent measurements of multiferroic MMn7 O12 (M=Ca, Sr, Cd,
Pb). Phys. Rev. B. 93(15), 155127 (2016)
84. Kagomiya, I., Matsumoto, S., Kohn, K., Fukuda, Y., Shoubu, T., Kimura, H., Noda, Y., Ikeda,
N.: Lattice distortion at ferroelectric transition of YMn 2 O 5. Ferroelectrics. 286(1), 167–174
(2003)
85. Chapon, L.C., Blake, G.R., Gutmann, M.J., Park, S., Hur, N., Radaelli, P.G., Cheong, S.W.:
Structural anomalies and multiferroic behavior in magnetically frustrated TbMn2 O5 . Phys.
Rev. Lett. 93(17), 177402 (2004)
86. Lee, N., Vecchini, C., Choi, Y.J., Chapon, L.C., Bombardi, A., Radaelli, P.G., Cheong, S.W.:
Giant tunability of ferroelectric polarization in GdMn2 O5 . Phys. Rev. Lett. 110(13), 137203
(2013)
87. Yahia, G., Damay, F., Chattopadhyay, S., Balédent, V., Peng, W., Elkaim, E., Whitaker, M.,
Greenblatt, M., Lepetit, M.B., Foury-Leylekian, P.: Recognition of exchange striction as the
origin of magnetoelectric coupling in multiferroics. Phys. Rev. B. 95(18), 184112 (2017)
88. Choi, Y.J., Yi, H.T., Lee, S., Huang, Q., Kiryukhin, V., Cheong, S.W.: Ferroelectricity in an
ising chain magnet. Phys. Rev. Lett. 100(4), 047601 (2008)
89. Aoyama, T., Yamauchi, K., Iyama, A., Picozzi, S., Shimizu, K., Kimura, T.: Giant spin-driven
ferroelectric polarization in TbMnO3 under high pressure. Nat. Commun. 5, 4927 (2014)
90. Arima, T.-h.: Ferroelectricity induced by proper-screw type magnetic order. J. Phys. Soc. Jpn.
76(7), 073702 (2007)
91. Kimura, T., Lashley, J.C., Ramirez, A.P.: Inversion-symmetry breaking in the noncollinear
magnetic phase of the triangular-lattice antiferromagnet CuFeO2 . Phys. Rev. B. 73(22),
220401 (2006)
618 J.-M. Hu and L.-Q. Chen

92. Seki, S., Onose, Y., Tokura, Y.: Spin-driven ferroelectricity in triangular lattice antiferromag-
nets ACrO2 (A=Cu, Ag, Li, or Na). Phys. Rev. Lett. 101(6), 067204 (2008)
93. Murakawa, H., Onose, Y., Miyahara, S., Furukawa, N., Tokura, Y.: Ferroelectricity induced by
spin-dependent metal-ligand hybridization in Ba2 CoGe2 O7 . Phys. Rev. Lett. 105(13), 137202
(2010)
94. Gao, X.S., Chen, X.Y., Yin, J., Wu, J., Liu, Z.G., Wang, M.: Ferroelectric and dielectric
properties of ferroelectromagnet Pb(Fe1/2 Nb1/2 )O3 ceramics and thin films. J. Mater. Sci.
35(21), 5421–5425 (2000)
95. Howes, B., Pelizzone, M., Fischer, P., Tabares-munoz, C., Rivera, J.-P., Schmid, H.: Charac-
terisation of some magnetic and magnetoelectric properties of ferroelectric Pb(Fe1/2 Nb1/2 )O3 .
Ferroelectrics. 54(1), 317–320 (1984)
96. Wang, J., Neaton, J.B., Zheng, H., Nagarajan, V., Ogale, S.B., Liu, B., Viehland, D.,
Vaithyanathan, V., Schlom, D.G., Waghmare, U.V., Spaldin, N.A., Rabe, K.M., Wuttig, M.,
Ramesh, R.: Epitaxial BiFeO3 multiferroic thin film heterostructures. Science. 299(5613),
1719–1722 (2003)
97. Lebeugle, D., Colson, D., Forget, A., Viret, M.: Very large spontaneous electric polarization
in BiFeO3 single crystals at room temperature and its evolution under cycling fields. Appl.
Phys. Lett. 91(2), 022907–022903 (2007)
98. Yakel Jr., H.L., Koehler, W.C., Bertaut, E.F., Forrat, E.F.: On the crystal structure of the
manganese(III) trioxides of the heavy lanthanides and yttrium. Acta Crystallogr. 16(10), 957–
962 (1963)
99. Bertaut, E.F., Mercier, M.: Structure magnetique de MnYO3 . Phys. Lett. 5(1), 27–29 (1963)
100. Magome, E., Moriyoshi, C., Kuroiwa, Y., Masuno, A., Inoue, H.: Noncentrosymmetric
structure of LuFeO3 in metastable state. Jpn. J. Appl. Phys. 49(9S), 09ME06 (2010)
101. Disseler, S.M., Borchers, J.A., Brooks, C.M., Mundy, J.A., Moyer, J.A., Hillsberry, D.A.,
Thies, E.L., Tenne, D.A., Heron, J., Holtz, M.E., Clarkson, J.D., Stiehl, G.M., Schiffer, P.,
Muller, D.A., Schlom, D.G., Ratcliff, W.D.: Magnetic structure and ordering of multiferroic
hexagonal LuFeO3 . Phys. Rev. Lett. 114(21), 217602 (2015)
102. Cox, D.E., Eibschütz, M., Guggenheim, H.J., Holmes, L.: Neutron diffraction study of the
magnetic structure of BaNiF4 . J. Appl. Phys. 41(3), 943–945 (1970)
103. Kato, K., Iida, S., Yanai, K., Mizushima, K.: Ferrimagnetic ferroelectricity of Fe3 O4 . J. Magn.
Magn. Mater. 31, 783–784 (1983)
104. Alexe, M., Ziese, M., Hesse, D., Esquinazi, P., Yamauchi, K., Fukushima, T., Picozzi, S.,
Gösele, U.: Ferroelectric switching in multiferroic magnetite (Fe3 O4 ) thin films. Adv. Mater.
21(44), 4452–4455 (2009)
105. Ikeda, N., Ohsumi, H., Ohwada, K., Ishii, K., Inami, T., Kakurai, K., Murakami, Y., Yoshii,
K., Mori, S., Horibe, Y., Kito, H.: Ferroelectricity from iron valence ordering in the charge-
frustrated system LuFe2 O4 . Nature. 436(7054), 1136–1138 (2005)
106. Lopes, A.M.L., Araújo, J.P., Amaral, V.S., Correia, J.G., Tomioka, Y., Tokura, Y.: New phase
transition in the Pr1−x Cax MnO3 system: evidence for electrical polarization in charge ordered
manganites. Phys. Rev. Lett. 100(15), 155702 (2008)
107. van den Brink, J., Khomskii, D.I.: Multiferroicity due to charge ordering. J. Phys. Condens.
Matter. 20(43), 434217 (2008)
108. de Groot, J., Mueller, T., Rosenberg, R.A., Keavney, D.J., Islam, Z., Kim, J.W., Angst, M.:
Charge order in LuFe2 O4 : an unlikely route to ferroelectricity. Phys. Rev. Lett. 108(18),
187601 (2012)
109. Niermann, D., Waschkowski, F., de Groot, J., Angst, M., Hemberger, J.: Dielectric properties
of charge-ordered LuFe2 O4 revisited: the apparent influence of contacts. Phys. Rev. Lett.
109(1), 016405 (2012)
110. Lafuerza, S., García, J., Subías, G., Blasco, J., Conder, K., Pomjakushina, E.: Intrinsic
electrical properties of LuFe2 O4 . Phys. Rev. B. 88(8), 085130 (2013)
111. Mundy, J.A., Brooks, C.M., Holtz, M.E., Moyer, J.A., Das, H., Rébola, A.F., Heron, J.T.,
Clarkson, J.D., Disseler, S.M., Liu, Z., Farhan, A., Held, R., Hovden, R., Padgett, E., Mao, Q.,
Paik, H., Misra, R., Kourkoutis, L.F., Arenholz, E., Scholl, A., Borchers, J.A., Ratcliff, W.D.,
12 Magnetoelectrics and Multiferroics 619

Ramesh, R., Fennie, C.J., Schiffer, P., Muller, D.A., Schlom, D.G.: Atomically engineered
ferroic layers yield a room-temperature magnetoelectric multiferroic. Nature. 537(7621),
523–527 (2016)
112. Liu, M., Hoffman, J., Wang, J., Zhang, J., Nelson-Cheeseman, B., Bhattacharya, A.:
Non-volatile ferroelastic switching of the Verwey transition and resistivity of epitaxial
Fe3 O4 /PMN-PT (011). Sci. Rep. 3, 1876 (2013)
113. Eerenstein, W., Morrison, F.D., Dho, J., Blamire, M.G., Scott, J.F., Mathur, N.D.: Comment
on “Epitaxial BiFeO3 multiferroic thin film heterostructures”. Science. 307(5713), 1203
(2005)
114. Zavaliche, F., Yang, S.Y., Zhao, T., Chu, Y.H., Cruz, M.P., Eom, C.B., Ramesh, R.: Mul-
tiferroic BiFeO3 films: domain structure and polarization dynamics. Phase Transit. 79(12),
991–1017 (2006)
115. Chu, Y.H., Martin, L.W., Zhan, Q., Yang, P.L., Cruz, M.P., Lee, K., Barry, M., Yang,
S.Y., Ramesh, R.: Epitaxial multiferroic BiFeO3 thin films: progress and future directions.
Ferroelectrics. 354(1), 167–177 (2007)
116. Silva, J., Reyes, A., Esparza, H., Camacho, H., Fuentes, L.: BiFeO3 : a review on synthesis,
doping and crystal structure. Integr. Ferroelectr. 126(1), 47–59 (2011)
117. Heron, J.T., Schlom, D.G., Ramesh, R.: Electric field control of magnetism using BiFeO3 -
based heterostructures. Appl. Phys. Rev. 1(2), 021303 (2014)
118. Sando, D., Barthélémy, A., Bibes, M.: BiFeO3 epitaxial thin films and devices: past, present
and future. J. Phys. Condens. Matter. 26(47), 473201 (2014)
119. Yang, J.-C., He, Q., Yu, P., Chu, Y.-H.: BiFeO3 thin films: a playground for exploring electric-
field control of multifunctionalities. Annu. Rev. Mater. Res. 45(1), 249–275 (2015)
120. Wu, J., Fan, Z., Xiao, D., Zhu, J., Wang, J.: Multiferroic bismuth ferrite-based materials for
multifunctional applications: ceramic bulks, thin films and nanostructures. Prog. Mater. Sci.
84, 335–402 (2016)
121. Sosnowska, I., Neumaier, T.P., Steichele, E.: Spiral magnetic ordering in bismuth ferrite. J.
Phys. C: Solid State Phys. 15(23), 4835 (1982)
122. Sosnowska, I., Zvezdin, A.K.: Origin of the long period magnetic ordering in BiFeO3 . J.
Magn. Magn. Mater. 140, 167–168 (1995)
123. Lebeugle, D., Colson, D., Forget, A., Viret, M., Bataille, A.M., Gukasov, A.: Electric-field-
induced spin flop in BiFeO3 single crystals at room temperature. Phys. Rev. Lett. 100(22),
227602 (2008)
124. Ederer, C., Spaldin, N.A.: Weak ferromagnetism and magnetoelectric coupling in bismuth
ferrite. Phys. Rev. B. 71(6), 060401 (2005)
125. Dzialoshinskii, I.E.: Thermodynamic theory of weak ferromagnetism in antiferromagnetic
substances. Sov. Phys. JETP. 5(6), 1259–1272 (1957)
126. Moriya, T.: Anisotropic superexchange interaction and weak ferromagnetism. Phys. Rev.
120(1), 91–98 (1960)
127. Dmitrienko, V.E., Ovchinnikova, E.N., Collins, S.P., Nisbet, G., Beutier, G., Kvashnin, Y.O.,
Mazurenko, V.V., Lichtenstein, A.I., Katsnelson, M.I.: Measuring the Dzyaloshinskii-Moriya
interaction in a weak ferromagnet. Nat. Phys. 10(3), 202–206 (2014)
128. Lebeugle, D., Mougin, A., Viret, M., Colson, D., Ranno, L.: Electric field switching of the
magnetic anisotropy of a ferromagnetic layer exchange coupled to the multiferroic compound
BiFeO3 . Phys. Rev. Lett. 103(25), 257601–257604 (2009)
129. Lee, S., Ratcliff II, W., Cheong, S.-W., Kiryukhin, V.: Electric field control of the magnetic
state in BiFeO3 single crystals. Appl. Phys. Lett. 92(19), 192906 (2008)
130. Popov, Y.F., Zvezdin, A., Vorob’Ev, G., Kadomtseva, A., Murashev, V., Rakov, D.: Linear
magnetoelectric effect and phase transitions in bismuth ferrite BiFeO3 . JETP Lett. 57, 69
(1993)
131. Ruette, B., Zvyagin, S., Pyatakov, A.P., Bush, A., Li, J.F., Belotelov, V.I., Zvezdin, A.K.,
Viehland, D.: Magnetic-field-induced phase transition in BiFeO3 observed by high-field
electron spin resonance: cycloidal to homogeneous spin order. Phys. Rev. B. 69(6), 064114
(2004)
620 J.-M. Hu and L.-Q. Chen

132. Tokunaga, M., Azuma, M., Shimakawa, Y.: High-field study of strong magnetoelectric
coupling in single-domain crystals of BiFeO3 . J. Phys. Soc. Jpn. 79(6), 064713 (2010)
133. Bai, F., Wang, J., Wuttig, M., Li, J., Wang, N., Pyatakov, A.P., Zvezdin, A.K., Cross, L.E.,
Viehland, D.: Destruction of spin cycloid in (111)c -oriented BiFeO3 thin films by epitiaxial
constraint: enhanced polarization and release of latent magnetization. Appl. Phys. Lett. 86(3),
032511–032513 (2005)
134. Sando, D., Agbelele, A., Rahmedov, D., Liu, J., Rovillain, P., Toulouse, C., Infante, I.C.,
Pyatakov, A.P., Fusil, S., Jacquet, E., Carrétéro, C., Deranlot, C., Lisenkov, S., Wang, D., Le
Breton, J.M., Cazayous, M., Sacuto, A., Juraszek, J., Zvezdin, A.K., Bellaiche, L., Dkhil, B.,
Barthélémy, A., Bibes, M.: Crafting the magnonic and spintronic response of BiFeO3 films
by epitaxial strain. Nat. Mater. 12(7), 641–646 (2013)
135. Zhao, T., Scholl, A., Zavaliche, F., Lee, K., Barry, M., Doran, A., Cruz, M.P., Chu,
Y.H., Ederer, C., Spaldin, N.A., Das, R.R., Kim, D.M., Baek, S.H., Eom, C.B., Ramesh,
R.: Electrical control of antiferromagnetic domains in multiferroic BiFeO3 films at room
temperature. Nat. Mater. 5(10), 823–829 (2006)
136. Chu, Y.-H., Martin, L.W., Holcomb, M.B., Gajek, M., Han, S.-J., He, Q., Balke, N., Yang,
C.-H., Lee, D., Hu, W., Zhan, Q., Yang, P.-L., Fraile-Rodriguez, A., Scholl, A., Wang,
S.X., Ramesh, R.: Electric-field control of local ferromagnetism using a magnetoelectric
multiferroic. Nat. Mater. 7(6), 478–482 (2008)
137. Heron, J.T., Trassin, M., Ashraf, K., Gajek, M., He, Q., Yang, S.Y., Nikonov, D.E., Chu, Y.H.,
Salahuddin, S., Ramesh, R.: Electric-field-induced magnetization reversal in a ferromagnet-
multiferroic heterostructure. Phys. Rev. Lett. 107(21), 217202 (2011)
138. Heron, J.T., Bosse, J.L., He, Q., Gao, Y., Trassin, M., Ye, L., Clarkson, J.D., Wang, C., Liu, J.,
Salahuddin, S., Ralph, D.C., Schlom, D.G., Iniguez, J., Huey, B.D., Ramesh, R.: Deterministic
switching of ferromagnetism at room temperature using an electric field. Nature. 516(7531),
370–373 (2014)
139. Zhou, Z., Trassin, M., Gao, Y., Gao, Y., Qiu, D., Ashraf, K., Nan, T., Yang, X., Bowden,
S.R., Pierce, D.T., Stiles, M.D., Unguris, J., Liu, M., Howe, B.M., Brown, G.J., Salahuddin,
S., Ramesh, R., Sun, N.X.: Probing electric field control of magnetism using ferromagnetic
resonance. Nat. Commun. 6, 6082 (2015)
140. Bibes, M., Barthélémy, A.: Multiferroics: towards a magnetoelectric memory. Nat. Mater.
7(6), 425–426 (2008)
141. Fusil, S., Garcia, V., Barthélémy, A., Bibes, M.: Magnetoelectric devices for spintronics.
Annu. Rev. Mater. Res. 44(1), 91–116 (2014)
142. Popkov, A.F., Kulagin, N.E., Soloviov, S.V., Sukmanova, K.S., Gareeva, Z.V., Zvezdin, A.K.:
Cycloid manipulation by electric field in BiFeO3 : coupling between polarization, octahedral
rotation, and antiferromagnetic order. Phys. Rev. B. 92(14), 140414 (2015)
143. Xue, F., Li, L., Britson, J., Hong, Z., Heikes, C.A., Adamo, C., Schlom, D.G., Pan, X., Chen,
L.-Q.: Switching the curl of polarization vectors by an irrotational electric field. Phys. Rev.
B. 94(10), 100103 (2016)
144. Rovillain, P., de Sousa, R., Gallais, Y., Sacuto, A., Méasson, M.A., Colson, D., Forget,
A., Bibes, M., Barthélémy, A., Cazayous, M.: Electric-field control of spin waves at room
temperature in multiferroic BiFeO3 . Nat. Mater. 9(12), 975–979 (2010)
145. Kundys, B.: Photostrictive materials. Appl. Phys. Rev. 2(1), 011301 (2015)
146. Kittel, C.: Domain boundary motion in ferroelectric crystals and the dielectric constant at
high frequency. Phys. Rev. 83(2), 458–458 (1951)
147. Qi, T., Shin, Y.-H., Yeh, K.-L., Nelson, K.A., Rappe, A.M.: Collective coherent control:
synchronization of polarization in ferroelectric PbTiO3 by shaped THz fields. Phys. Rev. Lett.
102(24), 247603 (2009)
148. Wen, H., Chen, P., Cosgriff, M.P., Walko, D.A., Lee, J.H., Adamo, C., Schaller, R.D., Ihlefeld,
J.F., Dufresne, E.M., Schlom, D.G., Evans, P.G., Freeland, J.W., Li, Y.: Electronic origin of
ultrafast photoinduced strain in BiFeO3 . Phys. Rev. Lett. 110(3), 037601 (2013)
149. Daranciang, D., Highland, M.J., Wen, H., Young, S.M., Brandt, N.C., Hwang, H.Y., Vattilana,
M., Nicoul, M., Quirin, F., Goodfellow, J., Qi, T., Grinberg, I., Fritz, D.M., Cammarata,
12 Magnetoelectrics and Multiferroics 621

M., Zhu, D., Lemke, H.T., Walko, D.A., Dufresne, E.M., Li, Y., Larsson, J., Reis, D.A.,
Sokolowski-Tinten, K., Nelson, K.A., Rappe, A.M., Fuoss, P.H., Stephenson, G.B., Lin-
denberg, A.M.: Ultrafast photovoltaic response in ferroelectric nanolayers. Phys. Rev. Lett.
108(8), 087601 (2012)
150. Iurchuk, V., Schick, D., Bran, J., Colson, D., Forget, A., Halley, D., Koc, A., Reinhardt, M.,
Kwamen, C., Morley, N.A., Bargheer, M., Viret, M., Gumeniuk, R., Schmerber, G., Doudin,
B., Kundys, B.: Optical writing of magnetic properties by remanent photostriction. Phys. Rev.
Lett. 117(10), 107403 (2016)
151. Hu, J.-M., Duan, C.-G., Nan, C.-W., Chen, L.-Q.: Understanding and designing magne-
toelectric heterostructures guided by computation: progresses, remaining questions, and
perspectives. npj Comput. Mater. 3(1), 18 (2017)
152. Rivera, J.P.: A short review of the magnetoelectric effect and relatedexperimental techniques
on single phase (multi-) ferroics. Eur. Phys. J. B. 71(3), 299 (2009)
153. Nan, C.W., Bichurin, M.I., Dong, S.X., Viehland, D., Srinivasan, G.: Multiferroic magneto-
electric composites: historical perspective, status, and future directions. J. Appl. Phys. 103(3),
031101 (2008)
154. Wang, Y., Hu, J., Lin, Y., Nan, C.-W.: Multiferroic magnetoelectric composite nanostructures.
NPG Asia Mater. 2(2), 61–68 (2010)
155. Srinivasan, G.: Magnetoelectric composites. Annu. Rev. Mater. Res. 40(1), 153–178
(2010)
156. Vaz, C.A.F., Hoffman, J., Ahn, C.H., Ramesh, R.: Magnetoelectric coupling effects in
multiferroic complex oxide composite structures. Adv. Mater. 22(26–27), 2900–2918 (2010)
157. Ma, J., Hu, J., Li, Z., Nan, C.-W.: Recent progress in multiferroic magnetoelectric composites:
from bulk to thin films. Adv. Mater. 23(9), 1062–1087 (2011)
158. Sun, N.X., Srinivasan, G.: Voltage control of magnetism in multiferroic heterostructures and
devices. Spine. 02(03), 1240004 (2012)
159. Hu, J.-M., Chen, L.-Q., Nan, C.-W.: Multiferroic heterostructures integrating ferroelectric and
magnetic materials. Adv. Mater. 28(1), 15–39 (2016)
160. Velev, J.P., Jaswal, S.S., Tsymbal, E.Y.: Multi-ferroic and magnetoelectric materials and
interfaces. Phil. Trans. R. Soc. A. 369(1948), 3069–3097 (2011)
161. Yu, P., Chu, Y.H., Ramesh, R.: Oxide interfaces: pathways to novel phenomena. Mater. Today.
15(7–8), 320–327 (2012)
162. Bibes, M.: Nanoferronics is a winning combination. Nat. Mater. 11(5), 354–357 (2012)
163. Vaz, C.A.F., Urs, S.: Artificial multiferroic heterostructures. J. Mater. Chem. C. 1(41), 6731
(2013)
164. Hu, J.-M., Shu, L., Li, Z., Gao, Y., Shen, Y., Lin, Y.-H., Chen, L.-Q., Nan, C.-W.: Film size-
dependent voltage-modulated magnetism in multiferroic heterostructures. Phil. Trans. R. Soc.
A. 372(2009), 20120444 (2014)
165. Hu, J.-M., Nan, C.-W., Chen, L.-Q.: Perspective: voltage control of magnetization in
multiferroic heterostructures. Natl. Sci. Rev. (2019). https://doi.org/10.1093/nsr/nwz047
166. Vaz, C.A.F.: Electric field control of magnetism in multiferroic heterostructures. J. Phys.
Condens. Matter. 24(33), 333201 (2012)
167. Liu, M., Sun, N.X.: Voltage control of magnetism in multiferroic heterostructures. Phil. Trans.
R. Soc. A. 372(2009), 20120439 (2014)
168. Ramesh, R.: Electric field control of ferromagnetism using multi-ferroics: the bismuth ferrite
story. Phil. Trans. R. Soc. A. 372(2009), 20120437 (2014)
169. Matsukura, F., Tokura, Y., Ohno, H.: Control of magnetism by electric fields. Nat. Nanotech-
nol. 10(3), 209–220 (2015)
170. Taniyama, T.: Electric-field control of magnetism via strain transfer across ferromagnetic/fer-
roelectric interfaces. J. Phys. Condens. Matter. 27(50), 504001 (2015)
171. Wang, Y., Li, J., Viehland, D.: Magnetoelectrics for magnetic sensor applications: status,
challenges and perspectives. Mater. Today. 17(6), 269–275 (2014)
172. Nikonov, D.E., Young, I.A.: Benchmarking spintronic logic devices based on magnetoelectric
oxides. J. Mater. Res. 29(18), 2109–2115 (2014)
622 J.-M. Hu and L.-Q. Chen

173. Ortega, N., Ashok, K., Scott, J.F., Ram, S.K.: Multifunctional magnetoelectric materials for
device applications. J. Phys. Condens. Matter. 27(50), 504002 (2015)
174. Hu, J.-M., Nan, T., Sun, N.X., Chen, L.-Q.: Multiferroic magnetoelectric nanostructures for
novel device applications. MRS Bull. 40(09), 728–735 (2015)
175. Newnham, R.E., Skinner, D.P., Cross, L.E.: Connectivity and piezoelectric-pyroelectric
composites. MRS Bull. 13(5), 525–536 (1978)
176. Lage, E., Kirchhof, C., Hrkac, V., Kienle, L., Jahns, R., Knochel, R., Quandt, E., Meyners,
D.: Exchange biasing of magnetoelectric composites. Nat. Mater. 11(6), 523–529 (2012)
177. Hu, J.-M., Li, Z., Wang, J., Nan, C.W.: Electric-field control of strain-mediated magnetoelec-
tric random access memory. J. Appl. Phys. 107(9), 093912 (2010)
178. Pertsev, N.A., Kohlstedt, H.: Resistive switching via the converse magnetoelectric effect in
ferromagnetic multilayers on ferroelectric substrates. Nanotechnology. 21(47), 475202 (2010)
179. Hu, J.-M., Li, Z., Chen, L.-Q., Nan, C.-W.: High-density magnetoresistive random access
memory operating at ultralow voltage at room temperature. Nat. Commun. 2, 553 (2011)
180. Binek, C., Doudin, B.: Magnetoelectronics with magnetoelectrics. J. Phys. Condens. Matter.
17(2), L39 (2005)
181. Chen, X., Hochstrat, A., Borisov, P., Kleemann, W.: Magnetoelectric exchange bias systems
in spintronics. Appl. Phys. Lett. 89(20), 202508–202503 (2006)
182. Hu, J.-M., Li, Z., Lin, Y.H., Nan, C.W.: A magnetoelectric logic gate. Phys. Status Solidi
RRL. 4(5–6), 106–108 (2010)
183. Hrkac, G., Dean, J., Allwood, D.A.: Nanowire spintronics for storage class memories and
logic. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 369(1948), 3214–3228 (2011)
184. Lei, N., Devolder, T., Agnus, G., Aubert, P., Daniel, L., Kim, J.-V., Zhao, W., Trypiniotis,
T., Cowburn, R.P., Chappert, C., Ravelosona, D., Lecoeur, P.: Strain-controlled magnetic
domain wall propagation in hybrid piezoelectric/ferromagnetic structures. Nat. Commun. 4,
1378 (2013)
185. Kanai, S., Yamanouchi, M., Ikeda, S., Nakatani, Y., Matsukura, F., Ohno, H.: Electric field-
induced magnetization reversal in a perpendicular-anisotropy CoFeB-MgO magnetic tunnel
junction. Appl. Phys. Lett. 101(12), 122403 (2012)
186. Wang, W.-G., Li, M., Hageman, S., Chien, C.L.: Electric-field-assisted switching in magnetic
tunnel junctions. Nat. Mater. 11(1), 64–68 (2012)
187. Shiota, Y., Nozaki, T., Bonell, F., Murakami, S., Shinjo, T., Suzuki, Y.: Induction of coherent
magnetization switching in a few atomic layers of FeCo using voltage pulses. Nat. Mater.
11(1), 39–43 (2012)
188. Grezes, C., Ebrahimi, F., Alzate, J.G., Cai, X., Katine, J.A., Langer, J., Ocker, B., Khalili
Amiri, P., Wang, K.L.: Ultra-low switching energy and scaling in electric-field-controlled
nanoscale magnetic tunnel junctions with high resistance-area product. Appl. Phys. Lett.
108(1), 012403 (2016)
189. Miwa, S., Suzuki, M., Tsujikawa, M., Matsuda, K., Nozaki, T., Tanaka, K., Tsukahara,
T., Nawaoka, K., Goto, M., Kotani, Y., Ohkubo, T., Bonell, F., Tamura, E., Hono, K.,
Nakamura, T., Shirai, M., Yuasa, S., Suzuki, Y.: Voltage controlled interfacial magnetism
through platinum orbits. Nat. Commun. 8, 15848 (2017)
190. Bauer, U., Emori, S., Beach, G.S.D.: Voltage-controlled domain wall traps in ferromagnetic
nanowires. Nat. Nanotechnol. 8(6), 411–416 (2013)
191. Bi, C., Liu, Y., Newhouse-Illige, T., Xu, M., Rosales, M., Freeland, J.W., Mryasov, O., Zhang,
S., te Velthuis, S.G.E., Wang, W.G.: Reversible control of co magnetism by voltage-induced
oxidation. Phys. Rev. Lett. 113(26), 267202 (2014)
192. Bauer, U., Yao, L., Tan, A.J., Agrawal, P., Emori, S., Tuller, H.L., van Dijken, S., Beach,
G.S.D.: Magneto-ionic control of interfacial magnetism. Nat. Mater. 14, 174–181 (2015)
193. Gilbert, D.A., Grutter, A.J., Arenholz, E., Liu, K., Kirby, B.J., Borchers, J.A., Maranville,
B.B.: Structural and magnetic depth profiles of magneto-ionic heterostructures beyond the
interface limit. Nat. Commun. 7, 12264 (2016)
194. Hu, J.-M., Yang, T., Wang, J., Huang, H., Zhang, J., Chen, L.-Q., Nan, C.-W.: Purely electric-
field-driven perpendicular magnetization reversal. Nano Lett. 15(1), 616–622 (2015)
12 Magnetoelectrics and Multiferroics 623

195. Wang, J.J., Hu, J.M., Ma, J., Zhang, J.X., Chen, L.Q., Nan, C.W.: Full 180◦ magnetization
reversal with electric fields. Sci. Rep. 4, 7507 (2014)
196. Hu, J.-M., Yang, T., Momeni, K., Cheng, X., Chen, L., Lei, S., Zhang, S., Trolier-McKinstry,
S., Gopalan, V., Carman, G.P., Nan, C.-W., Chen, L.-Q.: Fast magnetic domain-wall motion
in a ring-shaped nanowire driven by a voltage. Nano Lett. 16(4), 2341–2348 (2016)
197. He, X., Wang, Y., Wu, N., Caruso, A.N., Vescovo, E., Belashchenko, K.D., Dowben, P.A.,
Binek, C.: Robust isothermal electric control of exchange bias at room temperature. Nat.
Mater. 9(7), 579–585 (2010)

Hu is Assistant Professor of Materials Science and Engineer-


ing at University of Wisconsin (UW)-Madison. He received
his Ph.D. from Tsinghua University in Materials Science and
Engineering in 2013 and joined the faculty at UW-Madison in
2018. His current research activities focus on mesoscale compu-
tational modelling of ferroic materials and heterostructures and
computation-guided device designs based on these materials.

Chen is Hamer Professor of Materials Science and Engineering at


Penn State. He received his Ph.D. from MIT in Materials Science
and Engineering in 1990 and joined the faculty at Penn State
in 1992. He has published over 600 papers in the area of com-
putational microstructure evolution and multiscale modeling of
structural metallic alloys, functional oxides, and energy materials.
Magnetism and Superconductivity
13
Ilya M. Eremin, Johannes Knolle, and Roderich Moessner

Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 626
Paramagnetic Limit and Nonuniform FFLO Superconducting State . . . . . . . . . . . . . . . . . . . . 628
Interplay of Zeeman Field and Spin-Orbit Interaction: Spinless Fermions
and a Route Towards Topological Superconductivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 635
Ising Superconductors: Interplay of Magnetic Field and Spin-Triplet Channels . . . . . . . . . . . 638
Superconductivity in the Presence of Antiferromagnetic Order . . . . . . . . . . . . . . . . . . . . . . . . 640
Phenomenological Description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 640
Microscopic Consideration: Important Aspects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 643
Ferromagnetic Superconductors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 648
Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 651
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 652

Abstract

The interplay of magnetism and superconductivity is one of the most striking


features of the quantum mechanical description of solids. Originally thought to
be mutually exclusive, both phenomena in fact show interesting coexistence,
whose investigation remains a fascinating topic in contemporary condensed

I. M. Eremin ()
Institut für Theoretische Physik III, Ruhr-Universität Bochum, Bochum, Germany
e-mail: Ilya.Eremin@rub.de
J. Knolle
Blackett Laboratory, Imperial College London, London, UK
e-mail: j.knolle@imperial.ac.uk
R. Moessner
Max-Planck Institut für Physik komplexer Systeme, Dresden, Germany
e-mail: moessner@pks.mpg.de

© Springer Nature Switzerland AG 2021 625


J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_14
626 I. M. Eremin et al.

matter physics. In this short review, we have aimed to provide some introduction
to this exciting field of research.

Introduction

Magnetism has played an important role in our understanding of the collective


properties of matter for many decades, starting with the exact solutions of the
Ising model in one and two dimensions by Ising and Onsager, via the development
of the renormalization group all the way to the discovery of topological states of
matter in the form of spin liquids. Superconductivity has played a similarly central
role for the field of macroscopic quantum phenomena since its discovery at the
dawn of low-temperature physics by Kammerlingh Onnes and its explanation, much
later, in the BCS theory of Bardeen, Cooper, and Schrieffer. The latter being based
on a phononic mechanism, the two fields for a long time appeared to be at best
unconnected, and at worst mutually exclusive. While resolutions of this competition
started to be discussed already half a century ago, the discovery of high-temperature
superconductors thrust the relationship between superconductivity and a magnetism
at the very center of condensed matter physics 30 years ago.
Understanding this relationship remains a challenge to this day. The appearance
of long-range magnetic order is often associated with the localization of electrons,
whereas in the superconducting state “the opposite” happens as electrons are paired
and can then flow without resistance. Therefore, a priori one may think that these
two phenomena are mutually exclusive. Indeed, in bulk systems, the uniform spin-
singlet superconductivity and ferromagnetic order do not coexist. Vitaly Ginzburg
in 1957 demonstrated the problem of coexistence of magnetism and superconduc-
tivity by considering the interaction of a phenomenological superconducting order
parameter with a vector potential A of the magnetic field [1]. After the formulation
of the BCS theory in 1957 [2], it became also clear that superconductivity in
the spin-singlet state could also be destroyed by an exchange mechanism. In this
case, the exchange field in a magnetically ordered state tends to align spins of
Cooper pairs in the same direction, thus preventing a Cooper pairing. This is
the so-called paramagnetic effect [3]. Furthermore, Anderson and Suhl in 1959
demonstrated that ferromagnetic ordering is unlikely to appear in the spin-singlet
superconducting phase [4]. The main reason for this is the suppression of the
zero-wave-vector component of the electronic paramagnetic susceptibility (known
as the Pauli susceptibility) in a metal in the presence of superconductivity. In
such a situation, the energy for ferromagnetic ordering decreases and, instead of
ferromagnetic order, nonuniform magnetic ordering may appear, known by now
as cryptoferromagnetic. Furthermore, Larkin and Ovchinnikov and, independently,
Fulde and Ferrell showed that, in a pure ferromagnetic superconductor at low
temperature, superconductivity can only appear as a nonuniform solution [5, 6],
known by now as a LOFF (or FFLO) state.
The 1977 discovery of ternary rare-earth compounds ReRh4 B4 and ReMo6 X8
(here, Re = rare-earth element and X = S, Se) provided the first experimental
13 Magnetism and Superconductivity 627

evidence of magnetism and superconductivity coexisting in stoichiometric


compounds (see for a review [7]). It turned out that in these and in many
other systems, superconductivity coexists with antiferromagnetic order, and the
Neel temperature, TN , can be either larger or smaller than the superconducting
temperature, Tc , allowing to study their coexistence at both limits Tc >> TN and
Tc << TN as discussed in the corresponding section below. Since then, there
have been discoveries of further examples of systems where antiferromagnetism
and superconductivity coexist, including the quaternary intermetallic compounds
ReNi2 B2 C (see for a review [8]) and of course more recently heavy-fermion
superconductors (see for a recent review [9]), such as CeMIn5 (M = Co, Ir,
Rh) and iron-based superconductors [10, 11, 12]. Note that superconductivity
and antiferromagnetism can indeed coexist because the staggered magnetization,
associated with the AF state, appears to be zero at distances of the order of the
Cooper pair size or superconducting coherence length, ξ > 100 Å, or like in
unconventional superconductors the Cooper pairs do not feel the effective exchange
field due to the particular momentum dependence of the superconducting gap, as
happens in the extended s-wave or nodal d-wave gap in iron-based superconductors
and heavy-fermion superconductors, respectively.
Moreover, in iron-based superconductors, layered cuprates, and many heavy-
fermion systems, antiferromagnetism and superconductivity are strongly coupled,
even if they do not necessarily coexist. This indicates that the magnetic fluctuations
are also likely the driving force for the Cooper pairing instability in these systems.
One of the strong hints towards a common origin of antiferromagnetism and super-
conductivity in these compounds is the particular behavior of the superconducting
order parameter. The short-range antiferromagnetic fluctuations are usually peaked
at the characteristic antiferromagnetic wave vector, QAF . This provides a very
strong source of repulsion for the fermions, a priori inhibiting superconducting
Cooper pairing. To minimize this repulsion and to make it effectively attractive in
the Cooper channel within the BCS theory, the superconducting order parameter
Δ acquires a sign change for this wave vector, Δk = −Δk+QAF [13, 14, 15].
For the iron-based superconductors, the overall symmetry of the gap remains
anisotropic s-wave, while for the heavy-fermion superconductor CeMIn5 and
high-Tc cuprates, this condition yields dx 2 −y 2 -wave symmetry of the total order
parameter. Furthermore, in such a setting, a term appears in the free energy coupling
antiferromagnetic order and superconductivity, which can enhance the regions of
phase space where they coexist [16].
By contrast, the first ferromagnetic superconductors, UGe2 [17] and URhGe
[18], were discovered only relatively recently. These systems have spin-triplet
Cooper pairing, which allows the coexistence of superconducting order with
ferromagnetism more directly than in the singlet case. Indeed, superconductivity in
these systems appears below the Curie temperature of the ferromagnetic phase, Θ;
this makes the spin-singlet wave function of the Cooper pairs very improbable.
Nevertheless, the coexistence of singlet superconductivity with ferromagnetism
can be achieved in artificially fabricated layered ferromagnet/superconductor (F/S)
systems. Due to the proximity effect, the Cooper pairs can penetrate into the
628 I. M. Eremin et al.

ferromagnetic layer and induce superconductivity there. It is possible then to


study the interplay between superconductivity and ferromagnetism by varying the
layer thicknesses [19]. Note that the superconducting state appearing inside the
ferromagnet has a lot in common with the FFLO state including the oscillatory
behavior of the superconducting wave function.
Another interesting aspect of coexistence of superconductivity and magnetism is
the presence of the spin-orbit coupling, which breaks the spin-rotational invariance
of the system already in the paramagnetic state. This topic currently attracts a lot of
attention due to the phenomenon known as topological superconductivity. Although
it may formally also appear without magnetism, it represents an interesting example
of coexistence of spin-triplet (or more correctly odd-parity) Cooper pairing in the
presence of both spin-orbit coupling and the Zeeman exchange field [20, 21, 22].
This chapter is organized to give a simple and conceptual view on the interplay
of magnetism and superconducting order. After an outline of conventional Cooper
pairing, explaining the incompatibility of ferromagnetic and superconducting order,
we proceed with the examples of FFLO superconductors. Special emphasis is then
given to the microscopic coexistence of unconventional superconducting order and
antiferromagnetism. Here, we describe the coexistence of AF and SC states using
iron-pnictide superconductors as a prototypical example. Finally, we describe the
spin-triplet superconductivity in ferromagnetic superconductors.

Paramagnetic Limit and Nonuniform FFLO Superconducting


State

To gain an understanding of the paramagnetic effect, observe that the critical field
to destroy the superconducting state, Hp , at zero temperature may be found from a
comparison of two energy scales, the superconducting condensation energy, ΔEs ,
and the energy gain due to the spin polarization of electrons in the normal state,
ΔEp . The spin polarization of the electron gas changes its energy in the magnetic
field in the normal state by
H2
ΔEp = −χp , (1)
2

where χp = 2μ2B N(0) is the paramagnetic (Pauli) spin susceptibility of the normal
metal, μB is the Bohr magneton, N(0) is the density of electron states at the Fermi
level, and the electron g-factor is equal to 2. At the same time, the spin polarization
is absent in a spin-singlet superconductor, and the BCS condensation energy is given
by
Δ20
ΔEs = −N(0) , (2)
2

where Δ0 is the superconducting gap at zero temperature (in the BCS theory,
its value is given by 1.76kB Tc ). Comparing the two energy scales, one finds the
13 Magnetism and Superconductivity 629

paramagnetic limit at T = 0, known as the Chandrasekhar-Clogston limit [23, 24]:

Δ0
Hp (0) = √ . (3)
2μB

Note that this field would represent a first-order phase transition from a normal
to a superconducting state. However, Fulde and Ferrell [6] and Larkin and
Ovchinnikov [5] predicted the appearance of the nonuniform superconducting state
with a sinusoidal modulation of the superconducting order parameter at the scale
of the superconducting coherence length, ξs , which is now called FFLO state. In
this FFLO state, the Cooper pairs have a finite momentum compared with zero
momentum in conventional superconductors. The critical field of the second-order
transition into the FFLO state appears above the first-order transition line into
a uniform superconducting state [3]. At T = 0, it is H F F LO = 0.755Δ0 /μB ,
whereas Hp ∼ 0.7Δ0 /μB . The FFLO state only appears in the temperature interval
0 < T < T + (see Fig. 1) and is sensitive to impurities as it is a state with finite
momentum [25]. In real situations, the paramagnetic effect has to be considered in
addition to the orbital effects induced by the magnetic field. The relative importance
of the orbital and paramagnetic effects in the suppression of the superconductivity
is described by the Maki parameter [3]

√ Hc2
orb (0)
m ∗ Δ0
α= 2 ≈ , (4)
Hp (0) m0 εF

Fig. 1 (a) Schematic H − T phase diagram of three-dimensional spin-singlet superconductors


in the pure Pauli limit without orbital effect. Below T + = 0.56Tc , the transition becomes first
order [3]. The dashed line represents the metastable transition line.(b) H − T phase diagram in the
presence of the orbital effect. The transitions from the normal state to the FFLO and BCS states
are of second order (solid line), while the transition from the FFLO state to the BCS state is of first
order (thick solid line). The figure was reproduced with permission from [26]. Copyright 2007 The
Physical Society of Japan
630 I. M. Eremin et al.

where m∗ is the effective mass of the conduction electron and m0 is the free electron
mass. Gruenberg and Gunther examined the stability of the FFLO state against
the orbital effect for the spin-singlet three-dimensional superconductors [26]. The
FFLO state can exist at finite temperatures if α is larger than 1.8 and its region
narrows considerably from that in the absence of the orbital effect. The typical phase
diagram indicating the formation of the nonuniform FFLO state is shown in Fig. 1.
The appearance of a modulation in the superconducting order parameter is related
to the Zeeman splitting of the electron’s energy level under a magnetic field (or
exchange field if we are talking about ferromagnets) as illustrated in Fig. 2. In
the absence of the Zeeman field, a Cooper pair is formed by two electrons with
opposite momenta and spins, +kF ⇑ and −kF , ⇓. The resulting momentum of the
Cooper pair will be 0. Because of the Zeeman splitting, the Fermi momentum of the
electron with spin ↑ will shift from kF to k↑ = kF + δkF , where δkF = −μB H /vF
and vF is the Fermi velocity. Similarly, the Fermi momentum of an electron with
spin ↓ will shift from −kF to k↓ = −kF + δkF . The resulting momentum of
the Cooper pair will then be given by k↑ + k↓ = 2δkF , which implies that the
space modulation of the superconducting order parameter has a resulting wave
vector q = 2δkF . In contrast to one dimension, it is not possible to choose the single
wave vector δkF in two- [27] and three-dimensional superconductors [6, 5], which
compensates the Zeeman splitting for all electrons on the Fermi surface, as δkF
depends on the direction of vF , which yields the existence of the paramagnetic limit,
which is absent in one dimension. Observe that the critical field for a nonuniform
state at T = 0 is always higher than a uniform one. When T > μB H , at finite
temperature, the smearing of the electron distribution function near the Fermi energy
decreases the difference of energies between nonuniform and uniform states. From
microscopic calculations, at T > T + = 0.56Tc , the uniform superconducting phase

(a) (b) q
ky
ky

kx kx
Fig. 2 Schematic illustration of the pairing states in two dimensions. (a) BCS pairing state (k, ↑;
−k, ↓). (b) FFLO pairing state (k, ↑; −k + q, ↓) with q = 2δkF . The inner and outer circles
represent the Fermi surface of the spin-down and spin-up bands, respectively. The dashed circle
represents the initial Fermi surface
13 Magnetism and Superconductivity 631

is always favored [3]. Observe that in real systems the Maki parameter is usually
much less than unity; thus, the influence of the paramagnetic effect is negligibly
small in most superconductors. However, in quasi-2D layered superconductors (for
parallel fields) and heavy-fermion superconductors such as CeCoIn5 , αM is largely
enhanced owing to large m∗ /m0 values, and thus the superconductivity may be
limited by the Pauli paramagnetic effect. We discuss this in more detail below. It
should be stressed that in superconductors with small Fermi energies, a large Δ0 /εF
leads to the enhancement of αM , as was recently argued in some of the iron-based
superconductors such as KFe2 As2 [28] and FeSe [45].
The formation of the FFLO phase can be also described in the framework of the
generalized Ginzburg-Landau expansion [19]

b
F = a|ψ|2 + γ |∇ψ|2 + |ψ|4 , (5)
2

where ψ is the superconducting order parameter and the coefficient a is positive


for T > Tc and vanishes at T = Tc . Below Tc a is negative and the minimum in
the free energy functional F occurs for the uniform superconducting state |ψ|2 =
−a
b where b > 0. In a magnetic field, all coefficients depend on the energy of the
Zeeman splitting, μb H . In addition, the orbital effects of the magnetic field can be
included via gauge invariant substitution, ∇ → ∇ − (2ie/c)A. The FFLO state
occurs due to the fact that the coefficient γ changes its sign at the point (H + , T + )
of the phase diagram; see Fig. 1a. A negative sign of γ means that the minimum of
the functional does not correspond to a uniform state, and a spatial variation of the
order parameter decreases the energy of the system. To describe this properly, one
has to add a higher-order derivative term in the expansion (5), and the overall form
is now given by

η(H, T ) 2 2 b(H, T ) 4
F = a(H, T )|ψ|2 + γ (H, T )|∇ψ|2 + |∇ ψ| + |ψ| . (6)
2 2

Its minimization results in the equation that determines the critical temperature
of the second-order phase transition into a superconducting state

η 2
aψ − γ Δψ + Δ ψ = 0, (7)
2

where a = α(T − Tc (H )), where Tc (H ) is the critical temperature of the transition


into the uniform superconducting state. In the case of γ < 0, we search for the
nonuniform solution ψ = ψ0 eiqr . This yields the following equation:

η 4
a = −γ q 2 − q . (8)
2

Observe that for γ < 0, the maximum critical temperature corresponds to the
finite value of the modulation vector q02 = −γ /η and the corresponding transition
632 I. M. Eremin et al.

temperature
 F F LO into the nonuniform
 FFLO state TcF F LO (H ) with the coefficient a =
α Tc (H ) − Tc (H ) = γ /2η. This temperature is higher than the critical
2

temperature Tc (H ) of the uniform state. Thus, the appearance of an FFLO state


may simply be interpreted as a sign change of the gradient term in the Ginzburg-
Landau functional. Observe that a more detailed analysis of the FFLO state using
the generalized Ginzburg-Landau functional gives a sinusoidal modulation of the
order parameter [29, 30].
Formally, the generalized Ginzburg-Landau functional describes then a new type
of superconductor with very different properties, and the theory of superconductiv-
ity must be redone on the basis of this functional, and the detailed analysis can be
found in Ref. [19].
Concerning the experimental realization of the FFLO phase in the spin-singlet
superconductors, one has to bear in mind that there are several necessary conditions.
For example, this state may occur only in strongly type-II superconductors with very
large Ginzburg-Landau parameter, κ = λ/ξ , and large Maki parameter (here, λ is a
penetration depth and ξ is a superconducting coherence length), such that the upper
critical field approaches the paramagnetic limit. In addition, this superconductor has
to be in a clean limit, ξ << l, where l is a mean free path because the FFLO state
can be destroyed by nonmagnetic impurities. The main reason that the FFLO state
has never been observed in conventional superconductors resides in the fact that the
requirements for the formation of the FFLO state are extremely hard to fulfill.
The question of observing the FFLO state in unconventional spin-singlet super-
conductors has been addressed recently. There are some indications that the FFLO
state may exist in some organic superconductors and also CeCoIn5 . Yet in the
latter case, the situation is more subtle [31], and we will discuss it at a later
stage. Generally, heavy-fermion and organic superconductors are attractive classes
of superconductors for forming the FFLO state because of their potentially large
Maki parameters, αM ∼ 4 [31] (ξab ∼ 2 nm), very large mean free paths, and
superconducting pairing symmetries other than isotropic s-wave.
Next, we consider organic superconductors. These are materials that consist
of donor and acceptor molecules that form charge transfer salts. Among them
are the Bechgaard salts, (TMTSF)2 X with X = ClO4 , PF6 , and AsF6 , where the
tetramethyltetraselenafulvalene (TMTSF) molecules act as donors. These salts
contain a one-dimensional stacking of TMTSF molecules. They are conducting
because of π -orbital overlap between neighboring molecules. Superconductivity
in this class of materials was first found in (TMTSF)2 PF6 in 1979 [32]. There
are other important donor molecules for the synthesis of conducting organic
materials, including bis(ethylenedithio)- tetrathiafulvalene (BEDT-TTF) and
bis(ethylenedithio)tetraselenafulvalene (BETS). Crystalline structures built up of
BEDT-TTF form dimerized layered structures, leading to quasi-two-dimensional
electronic properties. A number of these materials also feature superconducting
phases at low temperatures, for example, for the acceptors Cu(NCS)2 [33],
Cu[N(CN)2 ]Cl [34], and Cu[N(CN)2 ]Br [35]. Among the organic superconductors,
there are two which indeed show the FFLO phase. In particular, κ-(BEDT-
TTF)2 Cu(NCS)2 is likely a spin-singlet d-wave superconductor [36]. It also features
13 Magnetism and Superconductivity 633

Fig. 3 H − T phase diagram


of κ-(BEDT-TTF)2 Cu(NCS)2
for fields in the
two-dimensional plane of the
dominant electronic
correlations, assembled from
differential susceptibility
measurements. Reproduced
with permission from [38].
Copyright 2000 IOP
Publishing Ltd. For further
discussion of experiments,
see also the recent review
by [42]

a rather unusual H − T phase diagram, with an upturn of Hc2 with decreasing


temperature (see Fig. 3). Pauli paramagnetic effects have been observed using 13 C
NMR [37]. For magnetic fields in the plane of the dominant electronic correlations,
the transition from the superconducting into the normal phase is of first order for
fields larger than 21 T, indicating the Pauli-limiting field [38, 39]. For higher fields,
an additional first-order transition appears inside the superconducting phase that was
interpreted as a transition to an FFLO phase [39]. Magnetic torque measurements
indeed confirm the presence of a high-field phase whose magnetic response is
distinctly different from that below the Pauli-limiting field [40]. More recently,
some unusual effects were observed in κ-(BEDT-TTF)2 Cu(NCS)2 using 1 3C NMR
measurements, which provide clear evidence of a change of the NMR spectra in the
low-field superconducting phase, and for superconductivity for fields higher than
the Pauli-limiting field. In particular, it was found that the NMR relaxation rate
increases in the putative FFLO phase, which was interpreted as evidence for polar-
ized Andreev bound states at the nodal regions of FFLO superconductivity [41].
In general, there should exist many more quasi-two-dimensional organic super-
conductors, which have large upper critical fields comparable to the paramagnetic
limit. However, experimentally, it is often challenging to observe these states. Some
of the potential candidates grow only as tiny crystals, so that specific heat or NMR
experiments mostly lack the necessary sensitivity. Nevertheless, there has been
recent reports of an FFLO state in the β −(ET)2 SF5 CH2 CF2 SO3 having the modest
bulk critical temperature of 4.3 K and the corresponding low Pauli-limiting field of
9.73 T [43]. Further experimental details on the study of this system can be found
in Ref. [42].
634 I. M. Eremin et al.

Finally, we note that some of the iron-based superconductors such as KFe2 As2
and FeSe can be an interesting platform for studying peculiar multiband features
of the FFLO phases, as both have Maki parameters larger than 1.85 and a very
large Ginzburg-Landau parameter, κ = 15, necessary conditions for the FFLO state
to be observed. In addition, in both systems, there are several bands crossing the
Fermi energy with relatively low EF . Especially in FeSe, an estimate gives Δ/εF ∼
0.1 − 0.3 for the hole band and an even larger Δ/εF for the electron band [44].
In KFe2 As2 , a clear double transition for the field strictly aligned in the ab-plane
and a characteristic upturn of the upper critical field line far beyond the Pauli limit
was observed in magnetotorque and specific heat measurements [28] as shown in
Fig. 4. In FeSe, it has also been discussed that the high-field phase for H ab can
be associated with an FFLO phase: in the H -T phase diagram shown in Fig. 4b, the
steep enhancement of Hc2 ab at low temperatures and the first-order phase transition

at a largely T -independent H ∗ are consistent with the original prediction of the


FFLO state [45]. At the same time, there are several unique aspects of the Fe-based
superconductors, which have to be taken into account in the original idea of the
FFLO state. In particular, the relatively small size of the Fermi surface pockets is
expected to give rise to the large spin imbalance near the upper critical field. In
addition, effective spin-orbit coupling, λso ∼ εF , yields a largely orbital-dependent
Zeeman effect and, finally, an orbital-selective pairing interaction is expected to
affect the q-wave vector modulation of the FFLO state. These call for a separate
detailed theory consideration.

Fig. 4 (a) H − T phase diagram of KFe2 As2 inferred from the magnetic torque and the specific
heat measurements; the data points are taken from Ref. [28]. The upper transition curve separates
the FFLO state from the normal paramagnetic (PM) state. The additional lines represent the Hc2
lines predicted for a purely Pauli-limited superconductor [46] and for an in-plane isotropic s-wave
and a d-wave superconductor with the FFLO state [47]. (b) High-field phase diagram of FeSe
for H (ab)-plane. Figure taken with permission from [45]. Copyright 2020 by the American
Physical Society. Blue circles and green crosses show Hirr and Hp determined by resistivity
measurements. Above the first-order phase transition field, a distinct field-induced superconducting
FFLO phase emerges at low temperatures
13 Magnetism and Superconductivity 635

Interplay of Zeeman Field and Spin-Orbit Interaction: Spinless


Fermions and a Route Towards Topological Superconductivity

Recent years have seen growing interest in topological superconductors, strongly


tied to the emergence of Majorana fermions, exotic particles which are their own
antiparticles [20, 48, 22]. In condensed matter systems, a quasiparticle which is
a superposition of electron and hole excitations so that its creation operator, γ † ,
becomes identical to the annihilation operator γ may be called as a Majorana
fermion. At zero energy, such Majorana modes are considered to be useful appli-
cation ingredients for topological quantum computation which can be manipulated
by braiding operations [49]. Although a review of topological superconductors is
well beyond the scope of this short chapter, we note here that Majorana fermions can
emerge in realistic experimental systems from the interplay of spin-orbit coupling
and a Zeeman magnetic field. In particular, two seminal works have established that
one can engineer a topological superconducting phase in a heterostructure consisting
of (i) a one-dimensional (1D) wire with appreciable spin-orbit coupling and (ii) a
conventional s-wave superconductor in a modest external magnetic field [50, 51].
The interface of the heterostructure can be modeled by the following Hamilto-
nian:

H = Hwire + HΔ , (9)

where
  
∂x2
Hwire = dx ψσ† − − μ − iασ ∂x + hσ ψσ
y z
(10)
2m

and

 
HΔ = dx Δ ψ↑ ψ↓ + h.c. . (11)

Here, ψσ† creates an electron with effective mass m, chemical potential μ, and
spin σ in the wire, while α > 0 denotes the strength of the spin-orbit coupling
and h ≥ 0 is the Zeeman energy arising from a magnetic field applied in the z-
direction. Overall, Hwire refers to an electron-doped semiconducting wire such as
InAs or InSb with strong Rashba spin-orbit coupling, in the limit where only the
lowest transverse subband is relevant. The pairing term HΔ models the proximity
effect arising from the adjacent s-wave superconductor like Nb or Al.
The red and blue curves in Fig. 5a illustrate the band structure of Hwire for h = 0.
Due to spin-orbit coupling, the blue and red parabolae respectively correspond to
electronic states whose spin aligns along y and −y. The magnetic field lifts the
crossing between these parabolae at k = 0, producing band energies
636 I. M. Eremin et al.

Fig. 5 (a) Electronic band structure for the interface between a wire and a conventional s-wave
superconductor, shown in the inset, without (red and blue curves) and with external magnetic field
(black curves). (b) Shows the kinetic energy in the Kitaev toy model with spinless fermions. The
superconducting gap opens a bulk gap except at the chemical potential values μ = ±t. For |μ| < t,
the system forms a topological weak-pairing phase. In this case, the model, written in terms of
Majorana fermions, couples at adjacent lattice sites, leaving two “unpaired” Majorana degrees of
freedom at the ends of the chain. Figures taken with permission from [20]. Copyright 2012 by IOP
Publishing Ltd.

k2 
± (k) = − μ ± (αk)2 + h2 (12)
2m

shown by the solid black curves in Fig. 5a.


When the Fermi level resides within this field-induced gap as shown in the
figure, the fermions in the wire appear to be effectively spinless. Focusing on
the spinless regime and assuming Δ << h, which allows to project away the
upper unoccupied band, one finds that Δ introduces p-wave pairs for the carriers,
driving the wire into a so-called topological superconducting state that connects
smoothlyto the Kitaev toy model featuring Majorana fermions [52]. In particular,
for h > μ2 + Δ2 (topological criterion), the Hamiltonian can be brought into the
simple form describing spinless fermions that hop on an N -site chain and exhibit
long-range-ordered p-wave superconductivity:

 1  †

HKitaev = −μ cx† cx − tcx cx+1 + Δe−iφ cx cx+1 + h.c. . (13)


2 x

Here, μ is again the chemical potential, t > 0 the nearest-neighbor hopping,


Δ the p-wave pairing amplitude, and φ the corresponding superconducting phase.
The chain bulk properties can be studied by imposing periodic boundary conditions
on the system. In momentum space, the Hamiltonian, HKitaev , can be brought to a
diagonal form with quasiparticle energy
13 Magnetism and Superconductivity 637


Ebulk (k) = εk2 + |Δ̃k |2 (14)

where εk = −t cos k − μ is the kinetic energy and Δ̃k = −iΔeiφ sin k is the pairing
amplitude in momentum space.
Ebulk (k) allows gapless bulk excitations when the chemical potential is fine-
tuned to μ = t or −t where the Fermi level respectively coincides with the top and
bottom of the conduction band. The gap closure at these isolated μ values reflects the
p-wave nature of the pairing required by the Pauli exclusion principle. In particular,
Δ̃k is an odd function of k, and Cooper pairing at k = 0 or k = ±π is prohibited,
thereby leaving the system gapless at the Fermi level when μ = ±t (see Fig. 5).
The physics of the chain is different in the two gapped regimes with |μ| > t
and |μ| < t. The first case connects smoothly to the trivial vacuum (upon taking
μ → ∞) where no fermions are present. By contrast, in the latter, a partially filled
band acquires a gap due to p-wave pairing. The new physics associated with this
state can be most simply accessed by decomposing the spinless fermion operators
cx in the original Kitaev Hamiltonian in terms of two Majorana fermions via

e−iφ/2  
cx = γB,x + iγA,x , (15)
2


where the γ operators obey the Majorana fermion relation γα,x = γα,x . In this basis,
the Hamiltonian, HKitaev , reads

μ 
N

HKitaev = − 1 + iγB,x γA,x
2
x=1


N −1
i
− (Δ + t) γB,x γA,x+1 + (Δ − t) γA,x γB,x+1 . (16)
4
x=1

To access the topologically nontrivial phase, one sets μ = 0 and t = Δ = 0.


Here, the Hamiltonian simplifies to

N −1
it 
HKitaev = − γB,x γA,x+1 (17)
2
x=1

which couples Majorana fermions only at adjacent lattice sites. However, as there
are two Majorana fermions per site, the ends of the chain (for the open boundary
conditions) now support “unpaired” zero-energy Majorana modes γ1 = γA,1 and
γ2 = γB,N . These can be combined into an ordinary – although highly nonlocal –
fermion that costs zero energy and therefore produces a twofold ground-state
degeneracy. Note the stark difference from conventional gapped superconductors,
638 I. M. Eremin et al.

where typically there exists a unique ground state with even parity so that all
electrons can form Cooper pairs.
The appearance of localized zero-energy Majorana end states and the associated
ground-state degeneracy arise because the chain forms a topological phase while the
vacuum bordering the chain is trivial. For the purpose of our review, it is important
that the fermions have to be spinless. This property ensures that a single zero-
energy Majorana mode resides at each end of the chain in its topological phase.
Suppose that instead spinful fermions, initially without spin-orbit interactions, form
a p-wave superconductor. In this case, spin merely doubles the degeneracy for
every eigenstate of the Hamiltonian, so that when |μ| < t, each end supports two
Majorana zero modes, or equivalently one ordinary fermionic zero mode. Unless
special symmetries are present, these ordinary fermionic states will move away from
zero energy upon including perturbations such as spin-orbit coupling.
Note that despite the intense search for intrinsic topological superconductors, the
practical realization of Majorana fermions appears to be easiest in the heterostruc-
ture described above with recent significant advances towards their experimental
realizations [53, 54, 55, 56, 57, 58]. This promises further exciting developments
on the interplay of magnetism and superconductivity in the field of topological
superconductors. To study the subject of topological superconductivity in depth,
the reader is invited to look into special reviews on the subject [20, 22, 53].

Ising Superconductors: Interplay of Magnetic Field and


Spin-Triplet Channels

In recent years, two-dimensional (2D) superconductivity became an active field of


research as a result of technological advances in the fabrication of heterostructures,
made from the van der Waals materials [59]. These systems are comprised of one-to-
several atomically thin monolayers exfoliated on substrates. Many of the properties
of the bulk persist down to the monolayer limit, especially in those systems where
the electronic structure is already quasi-two-dimensional one in the bulk samples.
For example, both bulk and monolayer NbSe2 are charge density wave metallic
superconductors [60,61]. However, due to the different effective dimensionality, the
monolayers react differently to the applied fields and often lack the inversion center,
present in the bulk system [62, 63]. These non-centrosymmetric superconducting
monolayers having in-plane mirror σh symmetry are called Ising superconductors.
Due to the time-reversal symmetry, the state |k ↑ (|k ↓) is degenerate with | − k ↓
(| − k ↑). When the lattice breaks parity, the spin-orbit coupling (SOC) causes
the spin splitting of Bloch states with the typical energy difference of Δso , and the
typical Fermi surface is shown in Fig. 6. As a result of the spin-orbit splitting, the
probability amplitude of the Cooper pair to be in a state |k ↑; −k ↓ differs from
the corresponding amplitude for the state |k ↓; −k ↑. Correspondingly, the parity-
even singlets and parity-odd triplets |Ψs,t  ∝ |k ↑; −k ↓ ∓ |k ↓; −k ↑ coexist
[64, 65].
13 Magnetism and Superconductivity 639

Fig. 6 Electronic structure of monolayer metallic transition-metal dichalcogenides as adapted


from Ref. [62]. Schematic representation of the crystal structure of 2Ha-MX2 viewed along [100]
direction for M = (Nb, Ta) and X = (S, Se) with 1H (monolayer). Spin-projected Fermi surface of
monolayer TaS2 and NbSe2 as computed by density functional theory (DFT) are also shown [62].
The variation of the shading and curve thickness refer to the magnitude of spin splitting in the
valence band due to spin-orbit coupling. Figures are taken from Ref. [62] using Creative Commons
CC BY license

Apart from inducing singlet-triplet mixing, the SOC makes the superconducting
state robust against the in-plane Zeeman field B. Because of the negligible thickness
of the monolayer, orbital-limiting effects do not contribute, and the only way
a magnetic field can affect the electronic states is via the paramagnetic effect,
discussed above. The large SOC enhances the critical in-plane field Bc well beyond
the Pauli limit. This has been studied theoretically [66, 65, 67, 68, 69, 70, 71] and
demonstrated experimentally [60, 72, 73, 61, 74, 75, 76, 62].
Till now, theoretical analyses of the superconducting pairing mechanism in
monolayer NbSe2 [71, 77] have relied on model descriptions of superconductivity
in materials that lack inversion symmetry based on the band structure calculated
from first principles. There is also a lack of consistency between first-principles
descriptions of superconductivity in bulk NbSe2 and experimental results. For exam-
ple, first-principles calculations overestimate Tc in bulk NbSe2 and isostructural
NbS2 and the zero-temperature gap [78]. Most recently, it was demonstrated that
bulk and monolayer NbSe2 are close to a magnetic instability [79], and repulsive
as well as spin-fluctuation-induced interactions cannot be most likely neglected
when addressing superconductivity in NbSe2 [80]. This perspective on the role of
magnetism in monolayer NbSe2 will also be crucial to un- derstand and control
the superconducting properties of monolayer NbSe2 in the presence of an external
magnetic field or with heterostructures between monolayer NbSe2 and magnetic
materials.
640 I. M. Eremin et al.

Superconductivity in the Presence of Antiferromagnetic Order

Phenomenological Description

Similar to the case of the FFLO state, the competition between superconductivity
and antiferromagnetism near their finite-temperature phase transitions can be
described in terms of a Ginzburg-Landau theory of coupled order parameters
without referring to microscopic details; see, for example, [16]. The free energy
functional in the spatially homogeneous case is given by

as bs δ am 2 bm 4
FAF +SC (ψ, M) = |ψ|2 + |ψ|4 + |ψ|2 M2 + M + M , (18)
2 4 2 2 4

where ψ and M refer to the superconducting (SC) and antiferromagnetic (AFM)


order parameter, respectively, and the integration over d-dimensional space is
implicitly assumed. As above, ψ is a complex order parameter, characterized by
an amplitude and a phase, while M is a three-component vector encoding magnetic
ordering. The quadratic coefficients am = am,0 (T − TN,0 ) and as = as,0 (T − Ts,0 )
change sign at TN,0 and Tc,0 , which denote the Neel and SC transition temperatures
without order-parameter competition. The leading term of the order-parameter
coupling is characterized by the coefficient δ, where the sign of δ > 0 reflects that
orders compete with each other. We consider the situation where the transitions for
δ = 0 are second order and the quartic coefficients bm and bs are positive. This free
energy functional accounts for various phase diagrams for competing and coexisting
AFM and SC orders.
In particular, consider TN,0 (p) and Tc,0 (p) vary as a function of a physical
parameter p, which typically is pressure, magnetic field, or doping (electron density
and/or disorder). In the case when both transitions meet at p = p∗ , i.e., T ∗ =
Tc,0 (p∗ ) = Tm,0 (p∗ ), one has a multicritical point (p∗ , T ∗ ) in the phase diagram.
In the vicinity of this multicritical point, a simultaneous expansion of the order
parameters is allowed. The mean-field analysis of Eq. (18) near (p ∗ , T ∗ ) depends
on whether δ 2 > bs bm or δ 2 < bs bm . We are interested in the situation when AFM
and SC orders compete (the physically more realistic situation) δ > 0 and one can
define a dimensionless quantity [16]

δ
g=√ −1 . (19)
bm bs

For g < 0 (0 < δ < bm bs ), (p∗ , T ∗ ) is a tetra-critical point where the two
second-order phase lines cross, leading to a regime in the phase diagram where
simultaneous AFM and SC orders coexist homogeneously. This is shown in Fig. 7a;
both phases compete but do not exclude √ each other.
On the other hand, if g > 0 (δ > bm bs ), the competition between two phases
is sufficiently strong such that both phases are separated by a first-order transition
that terminates at the bi-critical point (p∗ , T ∗ ). Notice that if the parameter p jumps
13 Magnetism and Superconductivity 641

Fig. 7 Schematic phase diagrams (p, T ) for competing AFM and SC orders, after [16]. Here,
p is a generic physical parameter and T is temperature. Solid and dashed lines denote second-
order or first-order phase transitions. For g < 0, there is a tetra-critical point and a region of
homogeneous coexistence (a). For g > 0, there is a bi-critical point, shown in (b) and (c). If p
changes discontinuously across the first-order transition, there will be a region of heterogeneous
coexistence for p1 < p < p2 (b), while its conjugate variable hp changes continuously and the
phase diagram has only one first-order line (c). Figure taken with permission from [16]. Copyright
2010 by the American Physical Society

discontinuously from p1 to p2 at the first-order transition, there is an intermediate


regime p1 < p < p2 of heterogeneous phase coexistence (see Fig. 7b). A sharp line
of first-order transitions occurs if one considers the phase diagram as a function of
hp , the variable that is thermodynamically conjugate to p (see Fig. 7c).
Note that critical fluctuations will change the universal exponents near the critical
temperatures and the slopes of the phase lines, but they do not change the generic
behavior shown in Fig. 7 (see [81, 82]).
Both order parameters can be finite simultaneously for g < 0, and this regime is
often referred to as coexistence of AF and SC, referring to microscopic coexistence
of SC and AFM orders. This should not be confused with phase coexistence
in the thermodynamic sense. The area in Fig. 7a below the tetra-critical point
642 I. M. Eremin et al.

is a single thermodynamic phase characterized by two order parameters that are


simultaneously finite. Similarly, the tetra-critical point is not a point where four
phases coexist, which would not be allowed by the Gibbs phase rule, but a point
where the system is in a single phase and both order parameters are infinitesimal
simultaneously. Below the bi-critical point, coexistence of thermodynamic phases
only occurs for p1 < p < p2 , where macroscopic AFM and SC regions occur
together in the sample. Usually, one uses the term homogeneous coexistence of
AFM and SC orders below the tetra-critical point to refer to coexisting order, and
heterogeneous coexistence below the bi-critical point to refer to coexistence of
phases.
From the Ginzburg-Landau expression, it is easy to obtain the temperature
dependence of the magnetic moment in the SC phase in the case of homogeneous
coexistence

am,0 bs (TN,0 − T ) + as,0 δ(T − Tc,0 )


M2 (T ) = √ . (20)
bm bs − δ

Without phase competition, both order parameters decrease as function of


increasing temperature, dM2 /dT 2 < 0 and dψ 2 /dT 2 < 0. However, once as,0 δ >
am,0 bs , one finds dM2 /dT 2 > 0 in the superconducting state. Thus, below Tc ,
the ordered moment decreases with decreasing temperature and the same condition
implies a back bending of the antiferromagnetic phase boundary upon entering the
superconducting state, a feature which is indeed often observed experimentally [16].
The considerations above are quite general and refer to conventional and uncon-
ventional superconducting states coexisting with antiferromagnetic order either of
itinerant or purely localized nature. Nevertheless, experimentally, both the AFM
and SC orders homogeneously coexist more often in the case of unconventional
superconductors. To understand this in detail requires a microscopic description.
However, one should bear in mind that in unconventional superconductors, the
superconducting gap often has opposite signs on the different sections of the Fermi
surface. In this case, the staggered magnetization, M, can be viewed as a mediator
of an intrinsic Josephson coupling between those regions. Specifically for the case
of the iron-based superconductors, which are believed to have opposite signs of the
superconducting gaps, ψ1 and ψ2 on the electron and hole Fermi surface pockets,
this Josephson interband coupling can be written as [16]

EJ ∝ M2 |ψ1 ||ψ2 | cos θ, (21)

where θ is a relative phase between the SC order parameters of the two Fermi
surface sheets. Once the phase difference is π (i.e., the superconducting gaps have
opposite phases), this term contributes to the reduction of the generally positive
value of δ in Eq. (18). Therefore, the situation of Fig. 7a appears more likely for
these unconventional superconductors.
13 Magnetism and Superconductivity 643

Microscopic Consideration: Important Aspects

The interplay between AFM and SC has been investigated in many contexts and
has a long history [83, 84, 85, 86, 87, 88, 89, 90, 91, 92, 93, 16, 94, 95, 96, 97, 98]. In
the following, we are not going to review these works in detail but concentrate on
exposing the most important aspects.
The first important aspect of superconductivity in the presence of translational
symmetry breaking AFM order is the folding of the electronic structure and
formation of new bands below TN , as illustrated in Fig. 8. The details depend on
the particular model under consideration, but the main features can be clarified for
the simplest case of the single-band model. In particular, consider the single-band
Hubbard Hamiltonian on a square lattice

 †
 
H =− ti,j ciσ cj σ + U ni↑ ni↓ − μ niσ , (22)
i,j,σ i i,σ


where ciσ creates an electron on site i with spin σ . The interaction term U denotes
the energy cost associated with having two electrons on the same site. In reciprocal
space, the Hamiltonian reads

 † U   † †
H = k ckσ ckσ + c c c−k+qσ ckσ , (23)
2N σ k σ −k +qσ
kσ k,k ,q

(a) (b) (c)

QAF
ky

kx kx kx

Fig. 8 Evolution of the Fermi surface topology for the commensurate AF state based on the
single-band model for layered cuprates. Due to the breaking of translational symmetry at the
antiferromagnetic momentum QAF M = (π, π ), the original energy dispersion, εk , in the
paramagnetic metallic state folds with the one shifted by the antiferromagnetic momentum
εk+QAFM . The original large Fermi surface (a) in the first Brillouin zone gets reconstructed as
an antiferromagnetic gap, W , opens on the “hot spots” εk = εk+QAFM (b). For intermediate W ,
the antiferromagnetic state is still metallic, and the residual Fermi surface consists of small hole
pockets, centered on (±π/2, ±π/2), and electron pockets, centered on (±π, 0), (0, ±π )
644 I. M. Eremin et al.

with

k = −2t[cos(kx ) + cos(ky )] − 4t cos(kx ) cos(ky ) − μ. (24)

The parameter −t is the energy gain corresponding to hopping between neigh-


boring sites, and −t denotes the energy gain by hopping to next-nearest-neighbor
sites. The doping level of the system is controlled by changing the chemical
potential μ. The interaction between two electrons is first treated in the Hartree-
Fock approximation giving rise to AF ordering of the spins. We therefore consider
the mean-field Hamiltonian

   σ W
c

† † k kσ
HSDW = (ckσ ck+Qσ ) , (25)
σ
σ W k+Q ck+Qσ
k

U  † †
where W = − N k [ck+Q↑ ck↑  − ck+Q↓ ck↓ ] is the AF order parameter.
Diagonalization of the mean-field Hamiltonian leads to the following energy
 ±
spectrum: Ek = k+ ± (k− )2 + W 2 , k± = k 2k+Q . The magnetic gap equation
α,β

is usually solved self-consistently for a given set of values of the hopping integrals
t = 1, t , the Coulomb repulsion U , and the doping.
The electronic structure in the AF state folds and yields the new Fermi surface
topology, illustrated in Fig. 8. The problem immediately becomes a two-band one
as there are now two types of fermionic bands, α and β. These give rise to electron
and hole pockets. In this regard, it is then more practical to consider the Cooper
pairing in terms of the α and β fermions, obtained after diagonalization of the
mean-field AF Hamiltonian. In particular, for the magnetization pointing along the
z-direction, the quasiparticle operators are related to the bare electron operators by
the transformation:

ckσ = uk αkσ + vk βkσ , (26)


ck+Qσ = sign(σ )[vk αkσ − uk βkσ ]. (27)

Using the new basis allows to account naturally for the umklapp Cooper pairing,
ck,↑ c−k−QAFM ,↓ , absent in the paramagnetic state, and the renormalization of
the chemical potential in the AF state, μ. In addition, one has to bear in mind that
the AF state breaks spin-rotational invariance, which yields parity (momentum in
the reduced BZ and its inversion) as the only good quantum number of the Cooper
pairs. The Cooper pairing has to be considered between quasiparticles residing
† † † †
in the same band, i.e., αkσ α−kσ  and βkσ β−kσ . In addition, one has to take
into account the possibility of anomalous pairings between fermions belonging to
† †
different pockets, i.e., αkσ β−kσ . The complete mean-field Hamiltonian includes
then the pair scattering processes between normal and anomalous gaps, as well as
scatterings between anomalous gaps. Such a pairing has been considered in the
context of the iron-based superconductors [96] and layered cuprates [98] as it
13 Magnetism and Superconductivity 645

couples linearly to the normal intraband gaps with a coupling constant proportional
to the SDW order parameter.
† †
In principle, anomalous pairs of the form αkσ β−kσ  involve fermions far from
the Fermi level for T < TN , since the two bands are gapped by the antiferromagnetic
β
gap |Ekα − Ek | ≥ 2W . Nevertheless, the anomalous gaps become sizeable due to
the coupling to the normal intraband pairs and exist only in the AFM background,
since the coupling between normal and anomalous gaps is proportional to W . A
detailed analysis [98] of the interaction Hamiltonian reveals, however, that even
† † † †
parity intraband gaps, i.e., αk↑ α−k↓ −αk↓ α−k↑ , couple to even parity anomalous
interband gaps. In the iron-pnictide study of Ref. [96], the reported anomalous
gap was dubbed spin triplet and coupled to an even parity singlet intraband gap.
However, one has to bear in mind that in this case the triplet gap is actually
of even parity, which is allowed due to the band index. Nevertheless, the spin-
singlet and spin-triplet gaps do coexist in the AFM background both having the
same parity. As the spin-rotational and the time-reversal symmetry for a given
sublattice is broken, one could even find the so-called s + it even parity state, as
shown in Fig. 9. However, to avoid confusion, we believe that it may be better to
classify the superconducting gaps by parity rather than by spin quantum numbers, as
spin-rotational symmetry (and to some extent time-reversal symmetry) is explicitly
broken, while the parity is conserved.
So far, the interactions yielding superconductivity have not been specified. They
are often considered on a phenomenological level. At the same time, the most
interesting situation is when the Cooper pairing arises from the same interaction
which drives the AFM transition. In the single-band Hubbard model, the source
of Cooper pairing is the spin and charge fluctuations. In particular, higher-order
interactions in U generate superconductivity on top of the AFM order through

Fig. 9 The spin structure of the gap function in an AFM background for the iron-based supercon-
ductor and the arrows refer to the relative phase of the intra- and interband order parameter: (a)
pure spin-singlet (s) even parity s +− -wave state, (b) pure spin-triplet (t) even parity s ++ -state, and
(c) s + it state with π/2 phase difference between s- and t-components as adopted from Ref. [96].
Figure taken with permission from [86]. Copyright 2014 by the American Physical Society
646 I. M. Eremin et al.

longitudinal and transverse spin fluctuations, following the original proposals of


Refs. [99,90]. Since U connects the bare electrons, the diagrammatics are performed
in terms of the bare electron Green’s functions, while the Cooper pairing is better
thought of to take place between the quasiparticles of the AF state. Transverse
and longitudinal spin fluctuations give rise to fundamentally different interactions.
Inspection of the interaction vertex formulated in real space [98] shows that the
charge and longitudinal interaction vertices give rise to no spin flips whereas the
transverse interaction does. In the latter channel, there are the gapless Goldstone
modes of the AFM phase, which give rise to a divergent interaction potential
between the bare electrons. However, when we consider pairing between the
quasiparticles of the AF state, this divergence is removed by the coherence factors
as noted in earlier works [100, 101]. In the case of pairing between opposite
spin electrons, spin flip processes are possible, and the effective interaction is
mediated both by longitudinal and transverse spin fluctuations. If pairing occurs
between same spin electrons, only longitudinal spin fluctuations contribute and the
pairing potential does not include the bare Coulomb repulsion U since this acts
only between opposite spin electrons. The interaction Hamiltonian is formulated in
terms of the SDW quasiparticles, and in line with earlier work [90, 100], one finds
the interactions in the longitudinal and transverse channel individually, with the
transverse part of the interaction stated as a spin flip vertex explicitly. The details of
this procedure was elaborated in Ref. [98].
The complete doping evolution of the three leading superconducting order
parameters, dx 2 −y 2 , g, and p -wave, was discussed in Ref. [98] (see Fig. 10). On
the hole-doped side, the near degeneracy of the dx 2 −y 2 and g-wave solutions, which
was observed very close to half filling, is split when hole doping is increased and
the dx 2 −y 2 solution becomes clearly dominant. In the case of electron doping, the
d-wave solution is strongest very close to half filling even though this is not where

Fig. 10 Phase diagram of coexisting unconventional superconductivity and antiferromagnetic


order, adapted from Ref. [98]. (a) Three leading superconducting instabilities as a function of
filling for the three largest eigenvalues of the linearized gap equation. The AFM region is shown
by the green area. (b) The critical temperature, Tc = 1.13c e−1/λ , with the energy cutoff set to
c = 0.25 t and t = 400 meV. The superconducting instability is dx 2 −y 2 at all fillings. Figure taken
with permission from [98]. Copyright 2016 by the American Physical Society
13 Magnetism and Superconductivity 647

the longitudinal and transverse pairing potentials achieve their maximum strengths.
The reason is that in the limit where electron pockets are small, the structure of
the intrapocket pairing potentials is purely attractive and this strongly supports a
d-wave solution. At critical electron doping for which W → 0, the Fermi arcs
just touch the magnetic zone boundary, and as a consequence, nesting by Q on the
paramagnetic side is rapidly weakened upon increased electron doping. There is a
jump in the magnitudes of the eigenvalues for the three leading gap symmetries in
the paramagnetic state. Their ordering, however, remains the same as in the SDW
phase, with the p solution, which in the paramagnetic phase takes the simpler form
[cos(kx ) − cos(ky )] sin(kx ), the least favorable. The jump in the eigenvalues is due
to the weak nesting properties of the Fermi surface on the paramagnetic side. In
fact, the paramagnetic Fermi surface, which is a hole pocket centered at (π, π ), is
roughly circular, thereby preventing nesting not only at Q but also at any q-vector.
As a result, spin-fluctuation-mediated superconductivity rapidly dies off.
It is interesting that the arguments on the stability of the dx 2 −y 2 -wave symmetry
on the AF background change if the Cooper pairing is mediated by an exchange
interaction which involves only nearest-neighbor sites, while the discussed approach
above deals with the on-site Coulomb repulsion and its renormalization in the
Cooper channel at higher orders. The similarity of the treatment of the coexistence
phase in Refs. [102] and [98] allows for a direct comparison of the interaction
Hamiltonians. In the simplified case of only hole pockets at the Fermi surface, the
interaction Hamiltonian reads generally

Hhole = Γ (k, k )βkσ
† †
β−kσ β−k σ βk σ , (28)
k,k σ

where in the t − J -like model without double occupancy constraint, employed in


Ref. [102], the effective interaction takes the form

J (k − k ) 2
Γt−J (k, k ) = − [m (k, k ) + l 2 (k, k )]
2
−J (k + k )[n2 (k, k ) + p2 (k, k )], (29)

with J (q) = J [cos(qx ) + cos(qy )] and J > 0. In the Hubbard model we have

+−
ΓHub (k, k ) = Γk,k
z
± 2Γk,k , (30)

with the longitudinal and transverse effective interactions. The lower sign of Eq. (30)
belongs to the triplet channel, and this is the source of an effective intrapocket
repulsion on the hole-doped side. On the contrary, the effective interaction as stated
in Eq. (29) gives rise to a purely attractive triplet potential for k and k residing
on the same hole pocket, and therefore the p-wave solution would indeed appear
to be the dominating instability. Note also that the model of Ref. [102] cannot be
considered strictly as a strong-coupling limit of the single-band Hubbard model
648 I. M. Eremin et al.

as it does not include a constraint for no double occupancies of the fermions


explicitly. This could play an important role as the original strong-coupling study of
unconventional superconductivity driven by the spin waves, studied within the t − J
model with the constraint of no double occupancies [103], does find the dx 2 −y 2 -
wave symmetry of the superconducting gap to be the only stable solution.
Finally, let us mention another interesting phase, which is probably a result of
the induced spin density wave (Q-phase) in the d−wave (nodal) superconductivity
in an external magnetic field, observed in CeCoIn5 for fields in the basal plane. This
phase exists only once the superconductivity is present and gets rapidly suppressed
when the field is turned away from this plane (see [31] for a review and references
therein). The SDW modulation wave vector in this phase was found to be pinned
to the lattice and points along the nodal direction. This was considered as evidence
that electronic nesting plays a role in the formation of the SDW order. Its origin in
CeCoIn5 is not fully understood at present, and several microscopic theories have
been developed [104, 105, 106, 107, 108, 109, 110, 111]. In the first approach, the
appearance of the magnetic phase appears as an instability towards an FFLO state
of modulated d-wave superconductivity at high fields close to the upper critical
field. Superconductivity in this FFLO state features a long-wavelength modulation.
Magnetism occurs in the FFLO nodal regions as Andreev bound states and is
thus a consequence of the modulated nature of superconductivity [104]. In the
other approaches, the magnetism arises for different reasons either as a result of
normal Fermi pockets around the nodal regions of the d-wave superconducting
gap [108, 109] or due to confined nature of the vortex structure [110]. In another
interpretation, the spin excitations in the dx 2 −y 2 superconductor condense at high
fields into the ground state, thereby creating a novel superconducting state [111].
The debate which of these approaches best describes the SDW phase in CeCoIn5
is still ongoing. The coexistence of magnetic and superconducting order parameters
has made it difficult to unambiguously identify the nature of the Q-phase and its
origin. On a phenomenological level, it is an important observation that the onset
of the Q-phase inside the superconducting phase occurs through a second-order
phase transition [112]. This is in contrast to many of the FFLO-based scenarios
that predict a first-order transition that separates a uniform d-wave phase from a
high-field modulated d-wave phase. Clearly, further analysis is called for.

Ferromagnetic Superconductors

Coexistence of superconductivity and ferromagnetism has recently been discovered


[17, 18, 113] in several uranium-based compounds, UGe2 , URhGe, and UCoGe.
This points towards an unconventional spin-triplet Cooper pairing mechanism in
these systems. In the first two compounds, the Curie temperature, TC , is higher
than their critical temperatures for superconductivity Tsc by more than an order
of magnitude, while in UCoGe, TC /Tsc = 4 at ambient pressure (see Fig. 11).
Also, the upper critical field at low temperatures greatly exceeds the paramagnetic
13 Magnetism and Superconductivity 649

Fig. 11 Temperature-pressure phase diagram of several ferromagnetic superconductors, (a)


UGe2 , (b) URhGe, and (c) UCoGe. FM, SC, and PM stand for ferromagnetic, superconducting, and
paramagnetic phases, respectively; the data points are taken from Ref. [114]. Originally published
in J. Phys. Soc. Jpn. 83, 061011 (2014). © 2014 The Author(s)

limit field in the first three compounds [114]. Further ferromagnetism does not
suppress superconductivity with triplet pairing, and no traces of spatial modulation
of magnetic moment directions on a scale smaller than the coherence length have
been revealed [18]. It is therefore natural to assume that these ferromagnetic
superconductors are triplet superconductors similar to superfluid phases of 3 He.
One has to keep in mind, however, that unlike liquid helium, which is a completely
isotropic neutral Fermi liquid, superconductivity in ferromagnetic superconductors
develops in strongly anisotropic ferromagnetic metals. In particular, all these
materials have an orthorhombic structure with the magnetic moment oriented either
along the a-axis like it is in UGe2 or along the c-axis in URhGe and UCoGe. The
magnetic moments are mostly concentrated around uranium ions, and at T → 0,
they are equal to 1.4μB in UGe2 [115], 0.4μB in URhGe [116], and 0.07μB
in UCoGe [117]. Although these values are much smaller than the moment per
uranium atom as deduced from the uniform susceptibilities above TC , this is still
not sufficient to treat uranium compounds as completely itinerant ferromagnets.
Instead, they are dual localized and itinerant ferromagnets. The interaction between
conduction electrons by means of spin waves in a system of localized moments
is considered to be the most plausible pairing mechanism. Models of this type have
been previously applied to the superconducting antiferromagnet UPd2 Al3 [118] and
to reentrant superconduc- tivity in the ferromagnetic URhGe [119]. In application
to the orthorhombic ferromagnets, the general structure of the order parameter,
dictated by symmetry, was proposed by [120, 121]. Subsequently, the microscopic
formulation based on the pairing interaction due to exchange of magnetization
fluctuations in orthorhombic ferromagnets with strong magnetic anisotropy was also
recently formulated [122, 123]. In the following, we quickly outline the details of
this consideration assuming Tsc < TC .
In the ferromagnetic state, the simplest model system consists of two bands,
ε↓ (k) and ε↑ (k), for the spin-down and spin-up bands, respectively. The spin-triplet
superconducting state arising in a ferromagnetic metal consists of spin-up, spin-
down, and zero-spin Cooper pairs described by the vector order parameter
650 I. M. Eremin et al.

Δ(k, r) = Δ↑ (k, r)| ↑↑ + Δ0 (k, r)(| ↑↓ + | ↓↑) + Δ↓ (k, r)| ↓↓, (31)

where Δ↑ (k, r), Δ0 (k, r), and Δ↓ (k, r) are the spin-up, zero-spin, and spin-down
amplitudes of the superconducting order parameter, depending on the Cooper
pair center of gravity coordinate r and the momentum k of pairing electrons.
Equivalently, one can write

Δ(k, r) = (d(k, r)σ )iσ y , (32)

with σ = (σ x , σ y , σ z ) being the Pauli matrices and the complex vector

1 
d(k, r) = −Δ↑ (k, r)(x̂ + i ŷ) + Δ↓ (k, r)(x̂ − i ŷ) + Δ0 (k, r)ẑ, (33)
2

and x̂, ŷ, and ẑ are the unit vectors along the corresponding coordinate axes.
Most of the ferromagnetic superconductors are ferromagnetic orthorhombic
crystals with a strong spin-orbit coupling fixing the spontaneous magnetization
along one of the symmetry axes chosen as the z-direction. Superconducting states
with different critical temperatures are described by the basis functions of the
different irreducible representations of the symmetry group of the normal state of the
crystal [124]. There are only two different representations, dubbed A and B of the
group GF M [120, 121]. The resulting vector order parameters are five-component
dA (k, r) and four-component dB (k, r). In particular, the symmetry considerations
[120] give the following form:


ΔA (k, r) = k̂x ηx↑ (r) + i k̂y ηy↑ (r)

ΔA (k, r) = k̂x ηx↓ (r) + i k̂y ηy↓ (r) (34)

Δ0A (k, r) = k̂z ηz0 (r)

and


ΔB (k, r) = k̂z ζz↑ (r)

ΔB (k, r) = k̂z ζz↓ (r) (35)
Δ0B (k, r) = k̂x ζx0 (r) + i k̂y ζy0 (r)

where k̂x , k̂y , and k̂z are the components of the unit momentum vector k̂. It is
important to note that the five-component order parameter of the A state and the
four-component order parameter of the B state include by symmetry the zero-spin
components. In other words, they are not equal-spin-pairing states consisting of
Cooper pairs with opposite spins, as one would expect naively for the ferromagnetic
state. This property originates from the spin-orbit coupling. However, although
13 Magnetism and Superconductivity 651

the pairing amplitudes for the zero-spin components can indeed arise due to the
spin-orbit terms in the ferromagnet gradient energy, they were shown on the basis
of microscopic considerations to be small [125]. Therefore, one can in the first
approximation ignore this amplitude Δ0 in the A and B states.
In this case, one deals with a two-band superconducting state similar to the
A2 state of superfluid 3 He [126]. Microscopically, the interaction between two
electrons driving the Cooper pairing is assumed to be due to the attraction of one
electron by the magnetic polarization cloud of the other. The pairing of electrons in
a ferromagnetic metal occurs in an anisotropic medium due to polarization of the
electron liquid and the localized moments.
This theory in application to the ferromagnetic superconductors was elaborated
in the context of ferromagnetic superconductors [123, 125]. It is important to
mention that the results of the microscopic considerations seem to agree with the
pure symmetry considerations, yet the experimental verification of the symmetry
of the superconducting order parameter is still absent. Part of the complication is
that in the ferromagnetic superconductor, there is an internal field Hint acting on
the electron charges even in the absence of an external field. The internal magnetic
field in all uranium ferromagnets is larger than the lower critical field Hc1 . Hence,
the Meissner state is absent and the superconductor is in an Abrikosov vortex state
with spatially inhomogeneous distributions of the order parameter and the internal
magnetic field, which complicates the analysis. Nevertheless, a comprehensive
state-of-the-art review on the subject can be found in [125].

Conclusions

The interplay of magnetism and superconductivity is one of the most striking


features of the quantum mechanical description of solids. Originally thought to be
mutually exclusive, both phenomena in fact show interesting coexistence, whose
investigation remains a fascinating topic in contemporary condensed matter physics.
In this short review, we have aimed to provide some introduction to this exciting
field of research.
We have first analyzed the FFLO state, which appears in the spin-singlet
superconducting state in the particular case when the Pauli-limiting field, Hp , is
lower than the expected upper critical field, Hc2 , determined from orbital effects. As
a result, there appears a modulation in the superconducting order parameter, which
is related to the Zeeman splitting of the electron’s energy level under a magnetic
field (or exchange field for ferromagnets) acting on electron spins. Experimentally,
such a state is seen in some organic superconductors and more recently in some
iron-based superconductors. We then analyzed the interplay between spin-orbit
coupling and an external, or an internal exchange, magnetic field acted in proximity
to conventional superconductors as a platform for the formation of topological
superconductivity with (localized) surface zero-energy excitations. Due to particle-
hole symmetry in a superconductor within the energy scale of the superconducting
gap, these bound states are described in terms of Majorana fermions. Another recent
652 I. M. Eremin et al.

material advance is provided by the two-dimensional superconductors with broken


inversion symmetry, which together with a sizeable spin-orbit coupling gives rise
to a mixture of spin-singlet and spin-triplet superconductivity. Although this effect
is known previously for bulk superconducting, its presence in these 2D materials
(known as Ising superconductivity) offers another interesting line of research, which
will be studied extensively in the future. We have then described the peculiarities of
coexistence between antiferromagnetic order and spin-singlet superconductivity. As
their coexistence is not restructured by any symmetry arguments, there are multiple
examples ranging from conventional to unconventional superconductors, where
both orders are adjacent on the phase diagram. Furthermore, there is a strong belief
that antiferromagnetism and spin-singlet superconductivity with sign changing gaps
on parts of the Fermi surface (denoted as unconventional superconductors, as the
mechanism of the Cooper pairing is not due to electron-phonon interaction) are
a result of the very same underlying magnetic interaction and therefore naturally
appear together in those systems. The examples are heavy-fermion systems, layered
cuprates, and iron-based superconductors. Finally, we have also given a brief
introduction to the field of ferromagnetic superconductors, which in contrast to
the FFLO state are uniform spin-triplet superconductors on a background of the
orthorhombic ferromagnetic state.
These fields of research are very dynamic, and the most recent progress in this
area should serve as both a theoretical and experimental guide in the search for novel
systems exhibiting cooperative behavior of magnetism and superconductivity.

References
1. Ginzburg, V.L.: Zh.Eksp. Teor. Fiz. 31, 202 (1956) [Sov. Phys. JETP 4, 153 (1957)]
2. Bardeen, J., Cooper, L.N., Schrieffer, J.R.: Phys. Rev. 106, 162 (1957)
3. Saint-James, D.D., Sarma, D., Thomas, E.J.: Type II Superconductivity. Pergamon, New York
(1969)
4. Anderson, P.W., Suhl, H.: Phys. Rev. 116, 898 (1959)
5. Larkin, A.I., Ovchinnikov, Y.N.: Zh. Eksp. Teor. Fiz. 47, 1136 (1964) [Sov. Phys. JETP 20,
762 (1965)]
6. Fulde, P., Ferrell, R.A.: Phys. Rev. 135, A550 (1964)
7. Maple, M.B., Fisher, O. (eds.): Superconductivity in Ternary Compounds II, Topics in Current
Physics. Springer, Berlin (1982)
8. Müller, K.-H., Narozhnyi, V.N.: Rep. Prog. Phys. 64, 943 (2001)
9. Pfleiderer, C.: Rev. Mod. Phys. 81, 1551 (2009)
10. Stewart, G.R.: Rev. Mod. Phys. 83, 1589 (2011)
11. Johnston, D.C.: Adv. Phys. 59, 803 (2010)
12. Hirschfeld, P.J., Korshunov, M.M., Mazin, I.I.: Rep. Prog. Phys. 74, 124508 (2011)
13. Mazin, I.I., Schmalian, J.: Phys. C 469, 614 (2009)
14. Scalapino, D.J.: Rev. Mod. Phys. 84, 1383 (2012)
15. Chubukov, A.V.: Annual Rev. Cond. Matt. Phys. 3, 57 (2012)
16. Fernandes, R.M., Schmalian, J.: Phys. Rev. B 82, 014521 (2010)
17. Saxena, S.S., Agarwal, P., Ahilan, K., Grosche, F.M., Haselwimmer, R.K.W., Steiner, M.J.,
Pugh, E., Walker, I.R., Julian, S.R., Monthoux, P., Lonzarich, G.G., Huxley, A., Sheikin, I.,
Braithwaite, D., Flouquet, J.: Nature (London) 406, 587 (200)
13 Magnetism and Superconductivity 653

18. Aoki, D., Huxley, A., Ressouche, E., Braithwaite, D., Flouquet, J., Brison, J.-P., Lhotel, E.,
Paulsen, C.: Nature (London) 413, 613 (2001)
19. Buzdin, A.I.: Rev. Mod. Phys. 77, 935 (2005)
20. Alicea, J.: Rep. Prog. Phys. 75, 076501 (2012)
21. Stanescu, T.D., Tewari, S.: J. Phys. Condens. Matter 25, 233201 (2013)
22. Sato, M., Ando, Y: Rep. Prog. Phys. 80, 076501 (2017)
23. Chandrasekhar, B.S.: Appl. Phys. Lett. 1, 7 (1962)
24. Clogston, A.M.: Phys. Rev. Lett. 9, 266 (1962)
25. Aslamazov, L.G.: Zh. Eksp. Teor. Fiz. 55, 1477 (1968) [Sov. Phys. JETP 28, 773 (1969)]
26. Matsuda, Y., Shimahara, H.: J. Phys. Soc. Jpn. 76, 051005 (2007)
27. Bulaevskii, L.N.: Zh. Eksp. Teor. Fiz. 64, 2241 (1973) [Sov. Phys. JETP 37, 1133 (1973)]
28. Cho, C.-w., Yang, J.H., Yuan, N.F.Q., Shen, J., Wolf, T., Lortz, R.: Phys. Rev. Lett. 119,
217002 (2017)
29. Buzdin, A.I., Kachkachi, H.: Phys. Lett. A 225, 341 (1999)
30. Houzet, M., Meurdesoif, Y., Coste, O., Buzdin, A.: Phys. C 316, 89 (1999)
31. Kenzelmann, M.: Rep. Prog. Phys. 80 034501 (2017)
32. Jerome, D., Mazaud, A., Ribault, M., Bechgaard, K.: J. Phys. Lett. 41 L95 (1980)
33. Urayama, H., Yamochi, H., Saito, G., Nozawa, K., Sugano, T., Kinoshita, M., Sato, S.,
Oshima, K., Kawamoto, A., Tanaka, J.: Chem. Lett. 17, 55 (1988)
34. Williams, J.M., Kini, A.M., Wang, H.H., Carlson, K.D., Geiser, U., Montgomery, L.K., Pyrka,
G.J., Watkins, D.M., Kommers, J.M.: Inorg. Chem. 29, 3272 (1990)
35. Kini, A.M., Geiser, U., Wang, H.H., Carlson, K.D., Williams, J.M., Kwok, W.K., Vandervoort,
K.G., Thompson, J.E., Stupka, D.L.A.: Inorg. Chem. 29 2555 (1990)
36. Taylor, O.J., Carrington, A., Schlueter, J.A.: Phys. Rev. Lett. 99, 057001 (2007)
37. Wright, J.A., Green, E., Kuhns, P., Reyes, A., Brooks, J., Schlueter, J., Kato, R., Yamamoto,
H., Kobayashi, M., Brown, S.E.: Phys. Rev. Lett. 107, 087002 (2001)
38. Singleton, J., Symington, J.A., Nam, M.-S., Ardavan, A., Kurmoo, M., Day, P.: J. Phys.
Condens. Matter 12, L641 (2000)
39. Lortz, R., Wang, Y., Demuer, A., Böttger, P.H.M., Bergk, B., Zwicknagl, G., Nakazawa, Y.,
Wosnitza, J.: Phys. Rev. Lett. 99, 187002 (2007)
40. Bergk, B., Demuer, A., Sheikin, I., Wang, Y., Wosnitza, J., Nakazawa, Y., Lortz, R.: Phys.
Rev. B 83 064506 (2011)
41. Mayaffre, H., Krämer, S., Horvatic, M., Berthier, C., Miyagawa, K., Kanoda, K., Mitrovic,
V.F.: Nat. Phys. 10, 928 (2014)
42. Wosnitza, J.: Ann. Phys. (Berlin) 2017, 1700282 (2017)
43. Beyer, R., Bergk, B., Yasin, S., Schlueter, J.A., Wosnitza, J.: Phys. Rev. Lett. 109, 027003
(2012)
44. Kasahara, S., Watashige, T., Hanaguri, T., Kohsaka, Y., Yamashita, T., Shimoyama, Y.,
Mizukami, Y., Endo, R., Ikeda, H., Aoyama, K., Terashima, T., Uji, S., Wolf, T., Löhneysen,
H.V., Shibauchi, T., Matsuda, Y.: Proc. Natl. Acad. Sci. USA 111, 16309–16313 (2014)
45. Kasahara, S., Sato, Y., Licciardello, S., Čulo, M., Arsenijević, S., Ottenbros, T., Tominaga, T.,
Böker, J., Eremin, I., Shibauchi, T., Wosnitza, J., Hussey, N.E., Matsuda, Y.: Phys. Rev. Lett.
24, 107001 (2020)
46. Gurevich, A.: Phys. Rev. B 82, 184504 (2010)
47. Jin, B.: Physica (Amsterdam) C 468, 2378 (2008)
48. Beenakker, C.W.J.: Annu. Rev. Condens. Matter Phys. 4 113 (2013)
49. Nayak, C., Simon, S.H., Stern, A., Freedman, M., Das Sarma, S.: Rev. Mod. Phys. 80, 1083
(2008)
50. Lutchyn, R.M., Sau, J.D., Das Sarma, S.: Phys. Rev. Lett. 105, 077001 (2010)
51. Oreg, Y., Refael, G., von Oppen, F.: Phys. Rev. Lett. 105, 177002 (2010)
52. Kitaev, A.Y.: Usp. Fiz. Nauk (Suppl.) 171 (10) (2001) [A.Y. Kitaev, Physics-Uspekhi 44, 131
(2001)]
53. Lutchyn, R.M., Bakkers, E.P.A.M., Kouwenhoven, L.P., Krogstrup, P., Marcus, C.M., Oreg,
Y.: Nat. Rev. Mater. 3, 52 (2018)
654 I. M. Eremin et al.

54. Zhang, H., Liu, C., Gazibegovic, S., et al.: Nature 556, 74 (2018)
55. Sestoft, J.E., Kanne, T., Norskov Gejl, A., von Soosten, M., Yodh, J.S., Sherman, D.,
Tarasinski, B., Wimmer, M., Johnson, E., Deng, M., Nygard, J., Sand Jespersen, T., Marcus,
C.M., Krogstrup, P.: Phys. Rev. Materials 2, 044202 (2018)
56. Fornieri, A., Whiticar, A.M., Setiawan, F., Marin, E.P., Drachmann, A.C.C., Keselman, A.,
Gronin, S., Thomas, C., Wang, T., Kallaher, R., Gardner, G.C., Berg, E., Manfra, M.J., Stern,
A., Marcus, C.M., Nichele, F.: Nature 569, 89 (2019)
57. Jäck, B., Xie, Y., Li, J., Jeon, S., Andrei Bernevig, B., Yazdani, A.: Science 364, 1255 (2019)
58. Manna, S., Wei, P., Xie, Y., Law, K.T., Lee, P.A., Moodera, J.S.: Proc. Nat. Acad. Sci. 117,
8775 (2020)
59. Gibertini, M., Koperski, M., Morpurgo, A.F., Novoselov, K.S.: Nature Nanotechnology 14,
408 (2019)
60. Ugeda, M.M., Bradley, A.J., Zhang, Y., Onishi, S., Chen, Y., Ruan, W., Ojeda-Aristizabal,
C., Ryu, H., Edmonds, M.T., Tsai, H.-Z., Riss, A., Mo, S.-K., Lee, D., Zettl, A., Hussain, Z.,
Shen, Z.-X., Crommie, M.F.: Nat. Phys. 12, 92 (2016)
61. Dvir, T., Massee, F., Attias, L., Khodas, M., Aprili, M., Quay, C.H.L., Steinberg, H.: Nat.
Commun. 9, 598 (2018)
62. de la Barrera, S.C., Sinko, M.R., Gopalan, D.P., Sivadas, N., Seyler, K.L., Watanabe, K.,
Taniguchi, T., Tsen, A.W., Xu, X., Xiao, D., Hunt, B.M.: Nat. Commun. 9, 1427 (2018)
63. Smidman, M., Salamon, M.B., Yuan, H.Q., Agterberg, D.F.: Rep. Prog. Phys. 80, 036501
(2017)
64. Gor’kov, L.P., Rashba, E.I.: Phys. Rev. Lett. 87, 037004(2001)
65. Frigeri, P.A., Agterberg, D.F., Koga, A., Sigrist, M.: Phys.Rev.Lett. 92, 097001 (2004)
66. Bulaevskii, L.N., Guseinov, A.A., Rusinov, A.I.: Zh. Eksp.Teor. Fiz. 71, 2356 (1976); [Sov.
Phys. JETP 44, 1243 (1976)]
67. Samokhin, K.V.: Phys. Rev. B 78, 224520 (2008)
68. Ilic, S., Meyer, J.S., Houzet, M.: Phys. Rev. Lett. 119, 117001 (2017)
69. Möckli, D., Khodas, M.: Phys. Rev. B 98, 144518 (2018)
70. Möckli, D., Khodas, M.: Phys. Rev. B 99, 180505(R) (2019)
71. Möckli, D., Khodas, M.: Phys. Rev. B 101, 014510 (2020)
72. Xi, X., Wang, Z., Zhao, W., Park, J.-H., Law, K.T., Berger, H., Forro, L., Shan, J., Mak, K.F.:
Nat. Phys. 12, 139 (2016)
73. Saito, Y., Nakamura, Y., Bahramy, M.S., Kohama, Y., Ye, J., Kasahara, Y., Nakagawa, Y.,
Onga, M., Tokunaga, M., Nojima, T., Yanase, Y., Iwasa, Y.: Nat. Phys. 12, 144 (2016)
74. Liu, Y., Wang, Z., Zhang, X., Liu, C., Liu, Y., Zhou, Z., Wang, J., Wang, Q., Liu, Y., Xi, C.,
Tian, M., Liu, H., Feng, J., Xie, X.C., Wang, J.: Phys. Rev. X 8, 021002 (2018)
75. Sohn, E., Xi, X., He, W.-Y., Jiang, S., Wang, Z., Kang, K., Park, J.-H., Berger, H., Forro, L.,
Law, K.T., Shan, J., Mak, K.F.: Nat. Mater. 17, 504 (2018)
76. Nakata, Y., Sugawara, K., Ichinokura, S., Okada, Y., Hitosugi, T., Koretsune, T., Ueno, K.,
Hasegawa, S., Takahashi, T., Sato, T.: npj 2D Mater. Appl. 2, 12 (2018)
77. He, W.-Y., Zhou, B.T., He, J.J., Yuan, N.F., Zhang, T., Law, K.: Commun. Phys. 1, 1 (2018)
78. Heil, C., Ponce, S., Lambert, H., Schlipf, M., Margine, E.R., Giustino, F.: Phys. Rev. Lett 119,
087003 (2017)
79. Wickramaratne, D., Khmelevskyi, S., Agterberg, D.F., Mazin, I.I.: arXiv:2005.05497
80. Shaffer, D., Kang, J., Burnell, F.J., Fernandes, R.M.: arXiv:1904.05470
81. Kosterlitz, J.M., Nelson, D.R., Fisher, M.E.: Phys. Rev. B 13, 412 (1976)
82. Aharony, A.: J. Stat. Phys. 110, 659 (2003)
83. Baltensperger, W., Strassler, S.: Phys. Kondens. Mater. 1, 20 (1963)
84. Bulaevskii, L.N., Rusinov, A.I., Kulic, M.: J. Low Temp. Phys. 39, 255 (1980)
85. Nass, M.J., Levin, K., Grest, G.S.: Phys. Rev. Lett. 46, 614 (1981)
86. Machida, K.: J. Phys. Soc. Jpn. 50, 2195 (1981)
87. Gulacsi, M., Gulacsi, Z.S.: Phys. Rev. B 33, 6147 (1986)
88. Gabovich, A.M., Shpigel, A.S.: J. Phys. F: Met. Phys. 14, 3031 (1984)
89. Kato, M., Machida, K.: Phys. Rev. B 37, 1510 (1988)
13 Magnetism and Superconductivity 655

90. Schrieffer, J.R., Wen, X.G., Zhang, S.C.: Phys. Rev. B 39, 11663 (1989)
91. Amici, A., Thalmeier, P., Fulde, P.: Phys. Rev. Lett. 84, 1800 (2000)
92. Das, T., Markiewicz, R.S., Bansil, A.: Phys. Rev. B 77, 134516 (2008)
93. Parker, D., Vavilov, M.G., Chubukov, A.V., Mazin, I.I.: Phys. Rev. B 80, 100508(R) (2009)
94. Vorontsov, A.B., Vavilov, M.G., Chubukov, A.V.: Phys. Rev. B 81, 174538 (2010)
95. Knolle, J., Eremin, I., Schmalian, J., Moessner, R.: Phys. Rev. B 84, 180510(R) (2011)
96. Hinojosa, A., Fernandes, R.M., Chubukov, A.V.: Phys. Rev. Lett. 113, 167001 (2014)
97. Schmiedt, J., Brydon, P.M.R., Timm, C.: Phys. Rev. B 89, 054515 (2014)
98. Roemer, A.T., Eremin, I., Hirschfeld, P.J., Andersen, B.M.: Phys. Rev. B 93, 174519 (2016)
99. Scalapino, D.J., Loh, E. Jr., Hirsch, J.E.: Phys. Rev. B 34, 8190 (1986)
100. Rowe, W., Eremin, I., Rømer, A.T., Andersen, B.M., Hirschfeld, P.J.: New J. Phys. 17, 023022
(2015)
101. Frenkel, D.M., Hanke, W.: Phys. Rev. B 42, 6711 (1990)
102. Lu, Y.-M., Xiang, T., Lee, D.-H.: Nat. Phys. 10, 634 (2014)
103. Kuchiev, M.Yu., Sushkov, O.P.: Phys. C 218, 197 (1993)
104. Yanase, Y., Sigrist, M.: J. Phys. Soc. Jpn. 78, 114715 (2009)
105. Yanase, Y., Sigrist, M.: J. Phys. Condens. Matter 23, 094219 (2011)
106. Yanase, Y., Sigrist, M.: J. Phys. Soc. Jpn. 80, SA005 (2011)
107. Yanase, Y., Sigrist, M.: J. Phys. Soc. Jpn. 80, 094702 (2011)
108. Kato, Y., Batista, C.D., Vekhter, I.: Phys. Rev. Lett. 107, 096401 (2011)
109. Kato, Y., Batista, C.D., Vekhter, I.: Phys. Rev. B 86, 174517 (2012)
110. Suzuki, K.M., Ichioka, M., Machida, K.: Phys. Rev. B 83, 140503 (2011)
111. Michal, V.P., Mineev, V.P.: Phys. Rev. B 84, 052508 (2011)
112. Tokiwa, Y., Bauer, E.D., Gegenwart, P.: Phys. Rev. Lett. 109, 116402 (2012)
113. Huy, N.T., Gasparini, A., de Nijs, D.E., Huang, Y., Klaasse, J.C.P., Gortenmulder, T., de
Visser, A., Hamann, A., Görlach, T., Löhneysen, H.V.: Phys. Rev. Lett. 99, 067006 (2007)
114. Aoki, D., Flouquet, J.: J. Phys. Soc. Jpn. 83 061011 (2014)
115. Kernavanois, N., Grenier, B., Huxley, A., Ressouche, E., Sanchez, J.P., Flouquet, J.: Phys.
Rev. B 64 174509 (2001)
116. Prokes, K., Gukasov, A., Takabatake, T., Fujita, T., Sechovsky, V.: Acta Phys. Polon. B 34,
1473 (2003)
117. Prokes, K., de Visser, A., Huang, Y.K., Fak, B., Ressouche, E.: Phys. Rev. B 81, 180407(R)
(2010)
118. McHale, P., Fulde, P., Thalmeier, P.: Phys. Rev. B 70, 014513 (2004)
119. Hattori, K., Tsunetsugu, H.: Phys. Rev. B 87 064501 (2013)
120. Mineev, V.P.: Phys. Rev. B 66, 134504 (2002)
121. Mineev, V.P., Champel, T.: Phys. Rev. B 69, 144521 (2004)
122. Mineev, V.P.: Phys. Rev. B 83, 064515 (2011)
123. Mineev, V.P.: Phys. Rev. B 90, 064506 (2014)
124. Sigrist, M., Ueda, K.: Rev. Mod. Phys. 63, 239 (1991)
125. Mineev, V.P.: Phys. Uspechi 60, 121 (2017)
126. Ambegaokar, V., Mermin, N.D.: Phys. Rev. Lett. 30, 81 (1973)
127. J. Dod, in The Dictionary of Substances and Their Effects, Royal Society of Chemistry.
(Available via DIALOG, 1999),
128. Ismer, J.-P., Eremin, I., Rossi, E., Morr, D.K., Blumberg, G.: Phys. Rev. Lett. 105, 037003
(2010)
129. Terashima, T., Kikugawa, N., Kiswandhi, A., Choi, E.-S., Brooks, J.S., Kasahara, S.,
Watashige, T., Ikeda, H., Shibauchi, T., Matsuda, Y., Wolf, T., Böhmer, A.E., Hardy, F.,
Meingast, C., Löhneysen, H.V., Uji, S.: Phys. Rev. B 90, 144517 (2014)
130. F. Ahn, I. Eremin, J. Knolle, V. B. Zabolotnyy, S. V. Borisenko, B. Büchner, and A. V.
Chubukov, Phys. Rev. B 89, 144513, (2014)
Part II
Magnetic Materials
Magnetism of the Elements
14
Plamen Stamenov

Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 660
The Magnetism of Iron, Cobalt and Nickel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 662
Band Structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 662
Magnetic Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 664
Thin Films of Fe, Co and Ni . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 664
Iron, Steels and Other Iron-Based Alloys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 664
Phase Diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 664
Manganese and Chromium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 667
Manganese . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 667
Chromium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 668
Spin Density Waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 669
Rare Earths . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 671
Magnetism of the Rare Earths . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 671
Magnetic Structures and Phase Transitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 677
Fermi-Level Spin Polarisation of the Magnetic Elements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 678
Spin Polarisation of the 3d Elements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 680
Spin Polarisation of the 4f Elements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 682
Oxygen . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 685
Molecular Magnetism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 685
Other Examples of Magnetic Order in the p- and d-Shell Elements . . . . . . . . . . . . . . . . . . 689
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 691

Abstract

This chapter presents the magnetic properties of the elements in relation to


their magnetic structure, with emphasis on those that order magnetically in

P. Stamenov ()
School of Physics and CRANN, Trinity College, University of Dublin, Dublin, Ireland
e-mail: stamenov.plamen@tcd.ie

© Springer Nature Switzerland AG 2021 659


J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_15
660 P. Stamenov

bulk form – iron, cobalt, nickel, manganese, chromium, most of the rare earths
and oxygen. All except oxygen are metals. The importance of spin polarisation
at the Fermi level is illustrated by the three room-temperature ferromagnets:
iron, cobalt and nickel. Manganese and chromium are atypical antiferromagnets;
α-Mn is a multi-sublattice antiferromagnet, β-Mn is a spin liquid, and Cr exhibits
an incommensurate spin density wave. A summary of the magnetism of the
rare earth elements emphasises the effects of the crystal field, their moments
and exchange integrals on their magnetic structures and phase transitions.
Some magnetoelastic effects are discussed, and a detailed account of the spin
polarisation of the heavy rare earths concludes the section. Finally, more exotic
forms of magnetism such as the molecular antiferromagnetism of oxygen and
defect-induced magnetism, exhibited by p and d-shell elements, such as carbon
and ruthenium, are presented.

Introduction

This is a short tale of the very foundations of material magnetism – the properties of
the magnetic elements. It is also a rather confined story; for as far as possible, it is
restricted to the properties of the essentially pure elements in their bulk form, rather
than their compounds and alloys. Crystal structure is a critical factor for determining
the magnetism of a solid. A glance at the magnetic periodic table in Fig. 1 reveals
that most of the elements possess a magnetic moment when they exist as free atoms.
The moment is determined by Hund’s rules. Only those elements with filled s, p, d,
or f shells or an electronic configuration such as p2 or d 4 , where L and S cancel,
have no moment to call their own. But everything changes when they enter the
community of a solid. Bonding with their neighbours often effaces all traces of their
former identity and eliminates their electrons with unpaired spin, be it covalent or
metallic. The moments survive to order ferromagnetically below a Curie point TC
or antiferromagnetically below a Néel point TN in just 16 of the 88 stable elements.
Only four of these order above room temperature, the ferromagnets (Fe, Co, Ni and
Gd) and the antiferromagnet (Cr), as vividly illustrated on Fig. 1. The importance of
crystal structure is underlined by the behaviour of different allotropes of the same
element; α- or γ -Fe, α- or β-Mn, β- or γ -Ce, for example, each have quite different
magnetic properties.
A common, yet striking comparison can be made of the magnetic and super-
conducting periodic Table 2. Despite both superconductivity and ferromagnetic
order being critical phenomena occurring below some well-defined temperature,
superconductivity is twice as common amongst the pure element, as can be seen
from an inspection of Fig. 2, where at least 32 pure elements have non-vanishing
superconducting Tc , albeit at temperatures that are much lower than their magnetic
counterparts. No element exhibits both types of order. This, of course, reflects
the quantum mechanical origins of these two phenomena. While in the case of
elemental superconductivity, the underpinning requirements are to have a metal,
which the vast majority of elements are (having a partially filled valence band) and
14 Magnetism of the Elements 661

Fig. 1 The magnetic periodic table. (After: [1])

Fig. 2 The periodic table of the superconducting transition temperatures. (After: www.
webelements.com)
662 P. Stamenov

strong-enough electron-phonon coupling – something satisfied best by ‘bad’ metals


that have relatively high room-temperature electrical resistivity. The requirements
for magnetism are substantially different - namely, to have a partially filled 3d or
4f orbitals, with very few exceptions, which will be discussed later.

The Magnetism of Iron, Cobalt and Nickel

The magnetism of the 3d metals is a subject of discussion in many well-established


textbooks [1]. The conventional approach is to start from a band picture of the two
bands most strongly present at the Fermi level – 3d and 4s. While the dispersions
of the 3d states are rather flat, with an effective mass components substantially
bigger than m0 and the corresponding partial density of states (PDOS) has peaks
at or near EF , the 4s states are substantially more mobile, with steeper dispersions
and effective masses of about or less than m0 . The two families of states hybridise
courtesy of the s-d exchange interactions, on the scale of couple of eV below and
above EF . The s band dispersion starts somewhat lower, about −5 to −4 eV below
the Fermi level, while the bandwidth and spin-splitting of the d-states are both of
the order of 1 eV.
Within the conventional Stoner picture for the occurrence of spontaneous
ordering in metallic systems, the energy gain due to the exchange interactions upon
polarising or spin-splitting the states at or near the Fermi level must exceed the
energy cost of pushing what is to become the majority spin band below EF . This is
traditionally expressed with the help of the Stoner criterion:

SD (EF ) > 1 (1)

where S is the Stoner exchange parameter, valued at ∼1 eV for the d-states.

Band Structures

It is instructive to look into the detail of the DOS for the three primary 3d
ferromagnets and analyse the differences that exist between them. In the case of iron,
while the overall difference of the spin-up and spin-down DOS, integrated from EF
down, and therefore the net available magnetic moment is the largest, the DOS in the
vicinity of EF is actually the smallest. Courtesy of the large exchange splitting of the
peak of the d states ∼2.8 eV, a lot of the available DOS for spin-down is unoccupied,
while the peak of the spin-up states is pushed to about 1 eV below EF . This results in
a reasonably high positive DOS polarisation P > 50%. The situation is different for
Co, as of both the different crystal symmetry (hcp), rather than bcc, and the different
total electron count. The exchange splitting of the peak of the d states is substantially
smaller or about 2 eV. The purely DOS spin polarisation is negative, P ∼ −70%.
14 Magnetism of the Elements 663

Out of the three primary elemental ferromagnets, Ni suffers the most of the high
occupancy of d states, at 3d 8 . This minimises further the exchange splitting of the
peak of the d states to approximately 1 eV and minimises further the net disbalance
of the spin-up and spin-down occupied DOS, resulting in the smallest magnitudes of
the net polarisation and Curie temperature in the tetrad. More surprising is the fact
that the purely DOS polarisation is rather large and negative at P ∼ −80%, as the
Fermi level is ‘pinned’ at the large peak of the d-states spin-down PDOS. This large
negative polarisation has been the source of much controversy in the early efforts
for matching theoretical predictions of the band structure with experiment (Fig. 3).

Fig. 3 Calculated density of


states for the three primary 3d
ferromagnetic metals.
(After: [2])
664 P. Stamenov

Table 1 Magnetic properties of Iron, Co and Ni


Density Lattice parameter Tc Ms K1 λs
Element Structure (kg/m3 ) (pm) (K) (MA/m) ( kJ/m3 ) (10−6 )
Fe BCC 7874 287 1044 1.71 48 −7
Co HCP 8836 251 1388 1.45 530 −62
Ni FCC 8902 352 628 0.49 −5 −34

Magnetic Properties

Thin Films of Fe, Co and Ni

While thin films of the pure elements are rarely used for spin electronic applications,
perhaps with the exception of epitaxial Fe, grown on MgO, the 3d random alloys and
ordered phases are critical building blocks for modern devices. The reader is referred
to  Chap. 16, “Metallic Magnetic Thin Films,” for details on these (Table 1).

Iron, Steels and Other Iron-Based Alloys

Phase Diagram

Soft Steels
Electrical Steel Silicon is a common ternary addition to low-carbon steel, which
helps to suppress and delay the α → γ transition from ferrite to austenite, normally
occurring about 912◦ C. As can be seen in the phase diagram in Fig. 5, the nominal
phase border is at 3.8 at. %, for otherwise pure iron. About 6 at. % of Si are usually
added, to account for leftover carbon, which substantially extends the austenitic
stability region, potentially all the way down to 727◦ C (see Fig. 4). The discovery
of the process is attributed to Robert Hadfield, around 1900, who had realised that
the compositions around 6 at % Si are sufficiently ductile to hot-roll to thin sheets.
The optimal thickness depends largely of the frequency band of application and the
skin depth, typically at main frequencies of 50–60 Hz. The largest amounts (about 7
B$) are produced with thickness of about 300 μm. There are additional gains as Si
lowers both the anisotropy and the magnetostriction in mild steel, while increasing
its electrical resistivity and corrosion resistance.
Goss Texture While in motor and generator applications isotropic microstructure is
appropriate, as the direction of the conducted flux can change substantially during
a cycle of the rotary machine (typically close to the ends of the pole pieces), in
transformer applications, the direction of the conducted flux stays locally largely
restrained onto a particular axis. This is, of course, the case away from zero
magnetic induction B, when changes in magnetic induction result in Eddy currents
being induced perpendicular to core lamination. Providing a crystallographic
texture and associated magnetocrystalline anisotropy can, indeed, be beneficial to
14 Magnetism of the Elements 665

Fig. 4 The carbon-iron phase diagram, at carbon concentrations below the cementite border.
(After: www.himikatus.ru/art and www.asminternational.com, Binary Alloy Phase Diagrams,
1986, Ed. T. B. Massalski, ISBN: 978-0-87170-403-0)

the minimisation of the core losses. The additional restriction provided by the
anisotropy, especially active at small B values, helps to keep the Eddy current
density low and eases the process of domain growth along the easy axis (in-plane).
Permalloy Permalloy is an all-purpose soft magnetic material that has been called
‘the fourth ferromagnetic element’. It is a partially ordered f cc alloy of Fe and
Ni, with approximate composition Fe20 Ni80 . The saturation polarisation is a little
more than 1 Tesla, but the alloy owes its success to the fortuitous cancellation of
both magnetostriction and intrinsic anisotropy at the same composition – a lucky
coincidence in a binary alloy; both change sign at a composition close to 80% Ni.
The soft magnetic properties are essentially impervious to strain. The soft magnetic
properties are improved by annealing and by small alloy additions of Cu and Mo, to
make alloys known as Supermalloy and MuMetal which can have permeabilities as
high as 105 and are very effective for magnetic shields.
Permendur The name permendur is reserved for the soft magnetic alloys of iron
and cobalt with compositions close to 50/50. It can contain up to 1% carbon
and undergoes order-disorder transition at 730◦ C, as illustrated in Fig. 6. Modern
compositions often have additional alloying elements, such as vanadium and silicon,
used to increase the machinability by increasing the eutectoid temperature and
666 P. Stamenov

Weight Percent Silicon


0 10 20 30 40 50 60 70 80 90 100
1700

1538°C
1500 L
1410°C 1414°C
1394°C

1300 23.5
19.5
Temperature °C

(γ Fe) 1212 1212°C 1220°C 1207°C


1203°C
29.8 67 73.5
3.8 β 50.8
1200 ζα
1100
1060°C
(α Fe) α2 982°C
965°C 70
28.2 η 937°C
70.5
900 912°C
825°C
770°C ε
α1 ζβ (Si)
700 Magnetic
Trans.

500
0 10 20 30 40 50 60 70 80 90 100
Fe Atomic Percent Silicon Si

Fig. 5 The iron-silicon phase diagram. (After: www.himikatus.ru/art and www.asminternational.


com, Binary Alloy Phase Diagrams, 1986, Ed. T. B. Massalski, ISBN: 978-0-87170-403-0)

improving the ductility. The magnetic transition temperature can be as high as


965◦ C for slightly iron-rich compositions. Permendur has the highest saturation
polarisation of all commercial soft magnetic alloys, with saturation induction in the
range of 2.3–2.4 Tesla. It finds applications in specific motors and generators and
rather importantly as the material of choice for the manufacturing of pole pieces of
laboratory electromagnets.

Hard Steels
Special Steels
Magnetic and Anti-magnetic Stainless Steel
Stainless steels are generally separated into three main classes, based on their
underlying crystalline structure – austenitic (f cc), ferritic (bcc) and martensitic
(bct). Out of these only the first one can be classified as essentially non-magnetic or
anti-magnetic. This is due to the austenitic structure being antiferromagnetic, with
an ordering temperature, which is strongly dependent on the amount and type of the
additive(s) used to stabilise the γ -phase, with Cr and Ni being the most commonly
used ones, in concentrations of up to 20%. The transitions generally occur at
14 Magnetism of the Elements 667

Weight Percent Iron


0 10 20 30 40 50 60 70 80 90 100
1700

L 1538°C
1500 1495°C
(δFe)
∼33%,1476°C
1394°C

1300 (αCo,γFe)
Temperature °C

1121°C
1100
Ma
Tra g. ∼55%,985°C
ns.

∼75 912°C
900 ∼27 Ma
(αFe) Tra g.
ns.
∼50%,730°C
770°C
700

α1

500
0 10 20 30 40 50 60 70 80 90 100
Co Atomic Percent Iron Fe

Fig. 6 The Cobalt-Iron phase diagram. (After: www.himikatus.ru/art and www.asminternational.


com, Binary Alloy Phase Diagrams, 1986, Ed. T. B. Massalski, ISBN: 978-0-87170-403-0)

cryogenic temperatures (5–50 K) and create significant problems for the application
of stainless steels in the construction of sensitive magnetometers and their use in
components designed for operation in high magnetic fields. The judicious choice of
composition and thermal treatment allows for the delay of both the Neel transition
and the transformation to the equilibrium martensitic or ferritic phase to be delayed,
with rather large amount of super-cooling, all the way down to He temperatures [3].

Manganese and Chromium

Manganese

Bulk manganese metal is a close to cubic antiferromagnet with rather complex


atomic and magnetic structures. Above its TN = 95 K, α-Mn adopts a cubic I 4̄3m
with 58 atoms per unit cell. Below the transition temperature, it distorts tetragonally
and adopts a non-collinear magnetic structure. The unusual set of properties
originates in the competition between the tendencies to maximise bond strength
and spin moment as dictated by Hund’s rules. Short Mn-Mn distances result in anti-
ferromagnetic exchange and low moments and generally tend to quench magnetism.
668 P. Stamenov

Fig. 7 The magnetic


structure of α-Mn.
(After: [4])

Crystal Structure
The crystal structure of α-Mn may be represented as a closed-packed intermetallic
compound, containing strongly magnetic (MnI) and (MnII), the weaker (MnIII), and
the even weaker (MnIV). Low-temperature magnetic moments are, respectively, 2.8,
1.8 and 0.5 μB . Atomic moments and exchange interactions in manganese and its
alloys are very sensitive to the Mn-Mn separation (Fig. 4).
Magnetic Structure
The non-collinear magnetic structure is a result of the weakly magnetic (MnIV)
being arranged on the triangular faces of polyhedra and exhibiting a frustrated anti-
ferromagnetic coupling, somewhat similar to that of the triangular antiferromagnets.
The other magnetic moments are rotated slightly from their otherwise collinear
antiferromagnetic orientation as a consequence of their finite interaction with the
(MnIV) [4] (Fig. 7).

Chromium

Crystal Structure
Paramagnetic chromium has a relatively simple bcc crystal structure, with an a =
2.884 Å. This is, however, altered below its TN = 311 K, by effects such as exchange
14 Magnetism of the Elements 669

striction. Since the details of the atomic order are in this case dependent on the spin
order, it is useful to understand better the underlying reasons for the existence of the
so-called charge density wave (CDW) state.
Magnetic Structure
Chromium is the prototypical 3d itinerant antiferromagnet. The underlying structure
is that of an alternating (001) spin planes, with the ones containing the corner atoms
opossing the ones passing through the body centring ones. Should this spin structure
have been perfectly commensurate, it would have been described by [5] (Fig. 8):

 comm = 2π [001] = 2π [001] = {0, 0, 1}


Q (2)
aCr Λ

This would imply that for the commensurate structure and in thermodynamic
equilibrium, three different domains would co-exist with the spins aligned along the
three main crystallographic directions. Reality is somewhat more complicated, with
an incommensurate linearly polarised spin density wave (I-SDW) structure, being
observed below the Néel point. This is a sinusoidal modulation of the magnetic
moment density of the type:

 
 (r ) = μ
μ  0 sin Q ± · r (3)

with μ0 being about 0.5 μB at helium temperatures. The incommensurate wave


vector is defined as (Fig. 8):

 
 ± = 2π 1 1 2π
Q − [001] = (1 − δ) (4)
aCr ΛSDW aCr

Both the commensurate and incommensurate states have been the subjects of
very detailed experimental and theoretical investigations, imploying a variety of
techniques, including neutron scattering and ultrasound attenuation, in order to
characterise both the order parameters and their influence on the electronic structure.
A detailed account of the state of investigations on elemental chromium [6] and its
alloys [7] is given by Fawcett and collaborators.

Spin Density Waves

The microscopic theory of the formation of the commensurate SDW has been
the subject of many debates, but the underlying principle seems to be related to
670 P. Stamenov

Fig. 8 The magnetic structure of α-Cr and a typical spin density wave. (After: [5])

the nesting properties of the Fermi surface, in this case [6]. The nesting implies
that the electron and hole Fermi surfaces are rather similar and can be essentially
 In a situation, similar to the one
superimposed by translation by a nesting vector Q.
appearing in some rare earths, this leads to anomalous electron-magnon scattering,
with both static and dynamic consequences. The SDW wave vector Q therefore is,

in most cases, very close to Q.
Commensurate SDWs
Highly defective Cr, resulting, for example, from a cold-working process, can
display rather high Neel transition (up to 475 K) to a commensurate AFM struc-
ture. This is attributed to the build-up of strain-induced dislocations [6]. It is
worth noting that in the commensurate AFM state, the PDOS of the d states
in chromium is not completely unpolarised. The small differences in the local
environment on the α and β sites should result in a rather small, yet potentially
observable Fermi level polarisation as indicated on Fig. 9. It probably comes to
a little surprise that some finite polarisation (P ∼ 5%) can be recovered in
experiments involving nano- and micro-indentation-based contact formation, such
as PCAR.
Incommensurate SDWs
14 Magnetism of the Elements 671

Fig. 9 The d-states DOS of Cr in the commensurate spin density wave regime. (After: [6])

Rare Earths

Magnetism of the Rare Earths

The properties of the rare earth elements in their metallic form are determined
primarily by the relative localisation of the 4f electrons, whose states resemble
closely the ones of the free atoms and ions, and forming the majority contribution to
the magnetic moments. The 5d and 6s states, on the other hand, delocalise and serve
a dual purpose, as quasi-free carriers and as mediators of the exchange interactions.
This separation, although not absolute, is a good starting point for the discussion
of the properties of the 4f states. Their interactions with the surroundings can be
classified into two classes – single-ion interactions, which are independent on each
672 P. Stamenov

crystal site i and contribute energy sums to the Hamiltonian, which are independent
on the different ionic residues, and two-ion interactions, which involve sums over
pairs of ions, indexed by i and j .
Crystal Field
The most important single-ion contribution is the effect of the crystal field. This
is effectively the influence of the charge distribution around an ionic site with
a particular local point symmetry, resulting in the large magnetic anisotropies,
characteristic of the rare earths, even in their metallic form. In first-order terms,
the crystal field contribution to the potential energy can be written as:

eρ(R)
vcf (r) = dR (5)
|r − R|

where ρ(R) is the charge density distribution created by the surrounding ionic
residues (both electronic and nuclear parts). If the overlap of the 4f electronic cloud
can be considered small, the solution of the Laplace’s equation may be written in
terms of an expansion over spherical harmonics of the type:


vcf (r) = Am l
l r Ylm (r ) (6)
lm

with the coefficients of the expansion Aml , given by the appropriate integrals of the
type:

m 4π ρ(R)
Am = (−1) dR l+1 Yl−m (R ) (7)
l
2l + 1 R

Therefore, the interaction is represented by a set of multipoles of the 4f


electronic distribution, interacting with the components of the electric field present
at the site. The situation gets somewhat more complicated, when the overlap
between the 4f distribution and the neighbouring charges is not small. Of course,
mathematically, vcf (r) can still be expanded into spherical harmonics, obeying
the right local point symmetry; however, the coefficients are given by the above
relatively simple integrals and are generally not scaling with r l , either. For the 4f
electrons, the multipole contributions with l > 6 vanish. This however does not
significantly simplify the task of evaluating the remaining terms. For the cases, when
the crystal field splitting remains small, when compared to the spin-orbit splitting,
one can take J to be a good quantum number and write the sums and integrals in
terms of the J operators.

 2l + 1 
Hcf = l αl r
Am l
Olm (Ji ) (8)

i lm

The Racah operators O lm (J) are defined via the spherical harmonics, projected
over Cartesian coordinates, by replacing r by J and symmetrising. Convenient tables
14 Magnetism of the Elements 673

of these are readily available in the literature. The convenience in using the Racah
operators, however, is not complete, as they transform as tensor operators and
behave under rotations like spherical harmonics. It is, therefore, rather customary
in the field to introduce the Stephens operators Olm (J), so that the crystal field part
of the Hamiltonian can be written as [8]:

Hcf = Blm Olm (Ji ) (9)
i lm

The first few of the Stephens operators can be expressed in terms of X ≡ J (J +1)
and J± ≡ Jx ± iJy as:

1 2 
O22 = J+ + J−2
2
1
O21 = (Jz Jx + Jx Jz )
2
O20 = 3Jz2 − X

Despite the apparent simplicity of the expression for the crystal field, in terms of
the Stephens operators, the coefficients Blm are still quite difficult to estimate with
an adequate accuracy, due to the non-spherical shapes of the charge distributions and
the existence of substantial charge redistribution and shielding effects. Two practical
approaches are very often deployed. One is to consider an adjustable magnitude
point-charge model for the charge distribution (compensating for the electronic
screening). The other is to treat BlM simply as experimental parameters, fixed based
on experimental data. For the rather popular hexagonal case, the crystal field can be
specified by fixing the values of four independent parameters:
⎡ ⎤
 
Hcf = i⎣ Bl0 Ol0 (Ji ) + B66 O66 (Ji )⎦ (10)
l=2,4,6

For effectively all rare earths, these parameters have been determined experi-
mentally, taking into consideration data available from a number of experimental
techniques, with the resulting multiplet splitting typically being in the range of 10-
ns to 100-ths of meV.

Magnetoelastic Effects
When the lattice is strained, both the crystal field and the two ion terms of the Hamil-
tonian are affected. This results in a substantial magnetoelastic coupling energy and,
courtesy of the modest elastic constants of the rare earth metals, to the possibility
of having large additional equilibrium strain and/or large magnetostriction. As the
strain scaling of the elastic energy is quadratic and the one of the magnetoelastic
energy is linear, balance situations may occur, indeed, for non-zero values of
equilibrium strain. Within the pure rare earths and their alloys is, indeed, where we
674 P. Stamenov

Table 2 Magnetic properties of the rare earths


Density Lattice parameter Tc,N μs K2 λs
Element Structure (kg/m3 ) (pm) (K) (μB ) (kJ/m3 ) (10−3 )
La HCP 6146 377 – – – –
Ce HCP 6689 368 13.7* 0.6 – –
Pr HCP 6773 367 0.05* 2.7 – –
Nd HCP 7008 366 19.9* 2.2 – –
Pm HCP 7264 365 – – – –
Sm RH 7520 363 106* 0.13 – –
Eu BCC 5244 458 90.4* 5.1 – –
Gd HCP 7901 363 293 7.63 – –
Tb HCP 8230 361 220 9.34 27 4.2
Dy HCP 8551 359 89 10.33 31 2.2
Ho HCP 8795 358 20 10.34 10 0.75
Er HCP 9066 356 20 9.1 −19 −0.25
Tm HCP 9321 354 32 7.14 – –
Yb FCC 6966 549 – – – –
Lu HCP 9841 351 – – – –
* the highest of multiple transition temperatures

find the largest magnetostriction coefficients; see, for example, Table 2. An example
of great practical importance is the alloy system Terfenol-D (Tbx Dy1−x Fe2 ) , for
x ∼ 0.3 , which exhibits magnetostriction coefficient λs > 2 · 10−3 . Before leaving
strain behind, it should be mentioned that the magnetoelastic effects can, indeed, be
considered to be strain-dependent renormalisations of the crystal field parameters.
There is, however, also a dynamic side to the coupling, which results in the crossing
and anti-crossing of the dispersion relations of magnons and phonons, in particular
high-symmetry directions, and can result in the formation of combined excitations,
sometime referred to as elasto-magnons.
Out of the two-ion couplings, the most important one to consider is the indirect
exchange, by courtesy of which the ions are coupled via the conduction electrons.
The general form of this interaction can be written as [8]:

 
2
Hsf (i) = − drI (r − R)Si · s(r) = − drHi (r) · μ(r) (11)
N

with N being the number of ions, s(r) is the conduction electron spin density and
the magnitude of the exchange integral I is determined by the geometric overlap
between the charge distributions of the 4f and the conduction states. The same
expression also defines an inhomogeneous effective magnetic field Hi (r) acting on
the conduction electron moment density μ(r) = 2μB s(r). Employing the standard
approach of the generalised susceptibility theory, a spin located at Ri contributes to
the moment μiα like:
14 Magnetism of the Elements 675


1   
μiα (r) = dr χαβ r − r Hiβ (r ) (12)
V
β

or in vector notation μi = χ̂ · Hi .
In order to try and exploit the periodicity of the crystal lattice, it is convenient
to define the forward and backward Fourier transforms of the generalised non-local
conduction electro-susceptibility as:

1
χ (q) = drχ (r)e−iq·r
V

V
χ (r) = drχ (q)eiq·r
(2π )3

which is valid within a scalar approximation and written this way for simplicity.
Moments and Exchange Integrals
The additional term in the Hamiltonian, corresponding to the f − f coupling
thorough the conduction electrons, is then simplified, defining I (q) to be the Fourier
transform of the exchange integral to:

1
Hff = − JS (ij )Si · Sj (13)
2
ij

taking the form of the familiar Heisenberg exchange, with JS (ij ) computed as:

1 
JS (ij ) = JS (q)eiq·(Ri −Rj ) (14)
N q

where JS (q) is effectively the projected exchange integral in momentum space:

V
JS (q) = |I (q)|2 χ (q) (15)
Nμ2B

Within the famous RKKY (Rudderman-Kittel-Kasuya-Yosida) approximation,


the non-local interaction between the conduction electron spins s(r) is replaced by
a local approximation as:
Hsf = −I0 δ(r − R)s · S (16)

which allows for an approximate estimation of the interaction H (ij ) as:

1 1 m∗  
H (ij ) ∼ Si · Sj 4
2kF Rij cos(2kF Rij ) − sin(2kF Rij ) (17)
4 (Rij ) h̄
676 P. Stamenov

Fig. 10 The indirect exchange interaction coupling strength, as deduced from neutron scattering
measurements of magnon dispersions in Pr. (After: [8] and [9])

where kF = q(EF ) is the Fermi wavevector for the conduction electrons, m∗ is the
conduction electron effective mass and Rij = Ri − Rj . The form of this interaction
is decaying oscillatory, with it covering typical distances of up to a couple of
nanometre.
The appearance of the so-called Kohn anomalies in J (q), due to the resonance
matrix elements for interaction of filled and empty states in the immediate vicinity of
the Fermi surface, separated by q gets Fourier transformed into Friedel oscillations
in J (R). The long periodicity of these is associated with the smallness of the
difference in energy and momenta of the above conduction states and is critical
for the interpretation of both the magnetic structures of the rare earths and their
magnetic and electronic excitations.
A rather famous example of the verification of this type of dependence is shown
on Fig. 10 and is based on the processing of neutron data for the low-temperature
magnon excitations in praseodymium [9].
Yet, more generally, for cases, where L is large, a large orbital component
of the moment is expected, which can lead to anisotropic contributions to the
effective Heisenberg exchange Hamiltonian. A more general form of the interaction
is therefore required, with different interaction constants coupling the different
effective dipole moments of the 4f electrons:

1
Hdd = − Jαβ (ij )Jiα Jjβ (18)
2
ij
14 Magnetism of the Elements 677

or in other words an anisotropic dipole-dipole coupling. The determination of all


relevant matrix components of the Jˆ tensor, without taking advantage of all possible
symmetry cancellations, is generally an insurmountable task.
The smallest in magnitude part of the Hamiltonian is due to interactions between
the f and s states with the nuclear spins I

Hhf = A Ii · Ji (19)
i
with the hyperfine coupling constant A being of the order of several micro-electron
volts, in most cases, and the interaction represented with an effective contact form.
While generally quite negligible, at sufficiently low temperatures, this term can
influence the magnetic ordering of the rare earths. A particular example is the
ground-state order in Pr.
Magnetic data from [8]. Anisotropy data from [10].

Magnetic Structures and Phase Transitions

The basic modes of the magnetic structures, which can be found in hexagonal
systems and specifically in the heavy rare earths are illustrated on Fig. 11.

Fig. 11 Basic magnetic


structures of the heavy rare
earths. The moments, within
each hexagonal layer, are
parallel to each other.
(After: [8])
678 P. Stamenov

Fermi-Level Spin Polarisation of the Magnetic Elements

As of the critical dependence of the functionality of spintronic devices on the Fermi


level spin polarisation, the problem of its quantification for an arbitrary material
has occupied both theorists and experimentalists for the best part of half a century.
Multiple definitions exist, which can be summarised, for the most commonly used
experimental methods, following reference [11]. The appropriate power n and the
nature of the energy and directional averaging of the corresponding velocities v
and density of states for the main quantisation axis D are different, depending
on the method of choice for the interrogation of the electronic system. The bare
DOS polarisation is almost impossible to access (the case of n = 0), although
some types of photoemission experiments do come close. The energy resolution
with which P (E) can be probed by means of photoemission is limited to about
100 meV, by the availability of high-resolution electron analysers and corresponding
flexible excitation synchrotron beam-lines. This limited resolution close to EF does
not usually permit direct comparisons with other methods. As a rule of thumb,
transport experiments, which depend on either spin-dependent tunnelling or quasi-
ballistic transport of electrons, would sample a form of the averaging for n ∼ 1,
with a relatively narrow interval of energies <1 eV around EF being probed.
These, unfortunately, require, in most cases, the construction of lithographically
defined structures and suffer from non-trivial dependences of the tunnelling matrix
elements on energy and spin direction. A stark example of structures, where
the matrix elements dominate the measured magnetoresistance changes and their
spectra, is the case of spin-filtering barriers, with Fe/MgO/Fe being the most
prominent prototype. In this limit, spin polarisation stops being a characteristic
of the electrode material but rather becomes a measure of the quality of coherent
specular transmission and reflection of electrons through a particular structure. Even
higher energy resolution(<1 meV) is available from methods relying on transport
within junctions, where one of the electrodes is superconducting. Two main
methods are well established: Tedrow-Messervey (TM) and Point Contact Andreev
Reflection (PCAR), which differ by the type of structures and superconductors
used.

D↑ v↑n − D↓ v↓n


P (E) = (20)
D↑ v↑n + D↓ v↓n

TM depends on the measurement and interpretation of the differential con-


dI
ductance spectrum dV (V ) of a tunnel junction of the ferromagnetic film being
studied, with a light (low spin-orbit coupling) superconductor, most commonly
Al. Relatively low temperatures are required, below the Tc of bulk aluminium
(<1.2 K), typically provided by 3 He or dilution refrigeration, in combination with
relatively high externally applied magnetic field ∼ T , sufficient to split the DOS
of the superconductor, beyond the experimental energy resolution. As the spin-
14 Magnetism of the Elements 679

split DOS of the superconductor produces antisymmetric effects in the junction


transport characteristic, it is possible to evaluate the sign of the spin polarisation,
by using TM. Some sensitivity towards the tunnelling matrix elements persists in
these measurements, however, which makes it difficult to evaluate the absolute value
of the spin polarisation |P | accurately, solely based on the assumption that the
conductance spectrum is dominated by the thermally and bias-voltage convoluted
singularities of the DOS within the superconductor. The appropriate value of n is
close to 1.
PCAR, on the other side, typically relies on dynamically formed point contacts,
using, in most cases, an indentor of superconducting niobium landing on a single
crystal or a polycrystalline surface of the material being studied. While variations
of PCAR exist, which allow for measurements at elevated temperatures (T <
30 K), for example, by using higher Tc superconductors, such as MgB2 , the vast
majority of measurement deployments of PCAR are either working at constant
4 He temperatures or are confined to about T < 12 K. The essence of the PCAR
dI
spectroscopic method is the measurement of dV (V ) of sufficiently transparent,
interfacial tunnelling barriers, which can be successfully modelled mathematically
by a Dirac delta function. This ensures that there is little energy dispersion of the
matrix elements for transport across the barrier and that the experimental spectra
are dominated by the bias dependence of the Andreev reflection process, itself. In
the simplest terms, the Andreev reflection process is one of the possible modes
for current conversion at the superconductor/normal metal interface, within which,
the conservation of energy, momentum and angular momentum is ensured, by the
‘reflection’ of one hole away from the interface into the normal metal for each
Cooper pair, incoming from the superconductor side. In the limiting cases of non-
magnetic metal and fully polarised half-metal, this process is correspondingly either
fully allowed or fully suppressed, as per the availability of final states of either
spin. PCAR spectra, in their simplest form, should therefore exhibit either doubling
of the normal-state junction conductance, within the bias window corresponding
the superconducting energy gap (for |P | = 1), or complete suppression thereof
(|P | = 0). The ratio of below gap conductance of the junction to the normal
state conductance (measured at sufficiently high bias |V | > kB Tc /e, or better yet
at sufficiently high temperature T > Tc ) is considered a direct measure of |P (EF |).
Unlike the case of TM, in PCAR the application of large external magnetic field
does not automatically lead to the determination of the sign of P . This limitation
is not fundamental, however, and specific cases exist (determined primarily by the
magnitude of the spin-orbit coupling within the ferromagnet), where it is possible
to resolve bias-anti-symmetric components of the PCAR spectra and define the sign
of P [12]. Generally, PCAR spectra can be analysed in both the ballistic n = 1 and
diffusive n = 2 limits. The determination of the transport regime for a particular
contact is problematic and required the measurement of spectra at different contact
pressures and temperatures, analysing in substantial detail the energy dispersions of
the spectral features, which are sensitive towards the type of velocity averaging at
EF . While the popular practice is to assume that sufficiently small point contacts
680 P. Stamenov

are intrinsically ballistic, as of the diminished probability for transverse momentum


scattering by phonons and magnons, the systematic differences induced in the
extracted values of |P | are less than a couple of percent and often below the
experimental uncertainties.

Spin Polarisation of the 3d Elements

The effects of the spin-orbit interaction in the transition metals are relatively week. It
is well-known that the latter is dominated by the crystal-field splitting, and generally
the orbital angular momentum is well-quenched. The existence of small orbital
momentum contributions, while experimentally verifiable, produces little effect on
the conduction electron spin polarisation [13]. The theoretical predictions for even
the most prototypical elemental ferromagnets vary substantially, depending on what
is the appropriate velocity averaging probed by a particular experiment. This is well
illustrated by the work of Mazin [11], which compares in detail the diffusive limit
spin polarisation for the magnetic 3d elements with the density of states one. The
DOS for the case of iron is featuring a rather high positive polarisation of about 70%,
while the same for Ni is even higher in magnitude, but opposite in sign: −80%,
close to EF . The latter is due to the essentially twice smaller exchange splitting
of the bands close to Fermi level and the band filling resulting in the equilibrium
Fermi level laying within the peak of the minority electron DOS. The situation is
altered dramatically in the ballistic case, with Fe and Ni both supposed to exhibit
|P | of between 35% and 45% and especially in the fully diffusive case, where Ni is
predicted to exhibit rather negligible polarisation (Fig. 12).
Spin-Polarised Photoemission
Before dabbing in the historical measurements of spin polarisation in the 3d
elements utilising photoemission, it is worth to note that some evidence of the
general understanding of the spin-split DOS can be obtained by experiments
involving optical excitation of direct intra- and inter-band transitions, too. This is
illustrated by a direct broad wavelength measurements of the optical conductivity
of single crystal specimen of iron, as illustrated on Fig. 13. The good agreement
between the calculated and computed absorption structure in the few eV vicinity of
EF justifies the validity of the band structure calculations based on local spin-density
approximation density functional theory.
The polarisation of the deeper states (10–100 eV) can be evaluated by techniques
related to photoemission, where electron beam rather than a photon one is bom-
barding the sample, such as Auger spectroscopy with polarisation analysis [14].
Polarisation levels as high as 70% are recovered, as illustrated in Fig. 14.
Later Photoemission Experiments
More elaborate experiments follow in UPS with and without polarisation of
the incoming photons and with and without polarisation analysis of the emitted
electrons. The latter analysis is performed, in most cases, using post-acceleration
14 Magnetism of the Elements 681

Fig. 12 Density of states and diffusion averaged density of states (dashed lines) for Fe (left panel)
and Ni (right panel) (After: [11])

to several keV and antisymmetric scattering from thin Au foil in a Mott detector.
These later investigations bring about more moderate values for the near-Fermi
level spin polarisation P of the 3d metals. One such example is the analysis
of thin in situ deposited films of Fe, Co and Ni by [16]. The reported values
of +54%, +21% and +15%, respectively, are all positive and shown to depend
little on the exact incoming photon energy and scale with the external field, in
a fashion resembling closely the bulk magnetisation of the materials. This latter
effect depends on two factors, the sensitivity towards projection of the photo-
electronic polarisation and the relatively large (millimetric) size of the illuminated
spots, containing a large number of magnetic domains. The exact relation between
the observed photoelectron polarisation and the DOS polarisation P (EF ) is, as
usual, more difficult to establish. The dominant contribution to the observed energy
distribution curves arises from a DOS peak, located at 0.30–0.35 eV below EF ,
attributed to the d-like states. The energy dispersion of the matrix elements to the
escape electron states has been shown to be negligible in the UPS of the transition
elements, as are the electron-magnon scattering, electron-plasmon coupling and
spin-orbit coupling. The sign problem, however, remains: most band structure
calculations predict negative band polarisation close to the Fermi level for Co and
Ni and positive for Fe; however, positive polarisation is evidenced experimentally
in both photo-emission and transport experiments.
682 P. Stamenov

σ⋅10–14, sec-1 σ, rel. un. σtot, rel. un.

60 b c 3
d 40
a
e
30
2 1
40 f
20
2

15

20 0

5 1

0
1 2 3 4 5 0 1 2 3 E, eV
E, eV

Fig. 13 Experimental optical conductivity in Fe and its separation into intra- and inter-band
components (left panel) and calculated total and spin-resolved optical conductivity (right panel).
(After: [15])

Field Emission
Considerable amount of experimental effort has been focused on the determination
of the spin polarisation of field-emitted electrons of both polycrystalline tips of Fe,
Co and Ni and material coated onto W tips. While wild variations are present in the
detected polarisation, depending on surface preparation, crystallographic direction
probed and adsorbed impurities, the maximal values quoted for Fe and Co are both
positive, at +80% and +47%, respectively; Ni shows a small negative polarisation of
−3.5% along the <111> axis, when pure and +15%, with hydrogen absorbed, which
makes for a difficult interpretation of the results [17].

Spin Polarisation of the 4f Elements

PCAR
Individual reports exist on the spin polarisation of the 4f elements by means of
PCAR. The spin polarisation in Gd and Dy has been evaluated at 52% and 50%,
respectively, for epitaxial thin films, grown on (1120) sapphire (Al2 O3 ). These
14 Magnetism of the Elements 683

Fig. 14 Raw experimental


data and corrected data for the
Auger electrons intensity and
polarisation. (After: [14])

are relatively high, especially when compared with the ones measured by other
superconductor-based techniques [18]. Thulium is another example of a heavy rare
earth, exhibiting relatively high P of 41%, even in a polycrystalline form [19].
Tedrow-Messervey
One of the most complete measurements of Fermi level polarisation in the heavy
rare earths has been performed using artificially structured Al-Al2 O3 -RE junctions
[20]. The same reveals tunnelling polarisations ranging from about 13% for Gd,
down to only 2.7% for Tm. Interestingly, within the restrictions of absolute accuracy
that the technique permits, an essentially linear scaling is established between the
net moment and the polarisation evaluated within about 1 meV of EF , as visualised
in Fig. 15. Unfortunately, this simple finding does not carry through the rest of
techniques for polarisation measurements, as will be discussed later.
SXPS, ARUPS and SARPES
The use of synchrotron radiation has allowed for faster and more precise work in
not only deep states spectroscopy but also UV photoemission spectroscopy, using
photons with both linear and circular polarisation. Apart from the determination
of surface states spin splittings as low as 10-ns of meV, these types of datasets also
684 P. Stamenov

Fig. 15 The scaling of the


polarisation of the heavy rare
earths, as seen by TM.
(After: [20])

allow for the extraction of the spin polarisation. An example is the case of Gd, where
a value of 64% has been determined [21].
Field Emission
Measurements of the spin polarisation of field-emitted electrons of rare-earth films
evaporated onto W tips [17] show large variability with preparation conditions,
with both positive and negative values ranging from −35% to +66%, often showing
considerable applied magnetic field dependence, as well. A focus on the positive
values, occurring more often and somewhat more stable towards small amounts of
surface contamination, reveals an even-odd pattern with the following values +53%,
+31%, +45%, +34%, + 66% and +13% for Gd, Tb, Dy, Ho, Er and Tm, respectively.
Polarisation of the Heavy Rare Earths
There have been rather few experimental investigations of the spin polarisation
of the heavy rare earths in the last couple of decades, despite the relative ease of
surface preparation for these, stemming directly from their better chemical stability.
While isolated efforts have been made to characterise individual elements in their
polycrystalline form, as the examples of Er and Tm in Refs. [12] and [19], are
useful for the illustration of discrepancies of the absolute values of P , evaluated
by different techniques and the illustration of the effects influence of spin-orbit
coupling in these, there is a growing need to reconcile the available data from
all accessible techniques and elements. Such an attempt is presented in Fig. 16. A
few peculiar observations can be made straight away. The available data segregates
into two groups of two techniques, which share large similarities in the trend of P
as a function of number of the element Z. The results of TM and photoemission
experiments reveal relatively low polarisation values, scaling almost linearly down
from about 15%, for Gd. Quite unlike the latter, the data from field emission
experiments and from PCAR exhibits much higher polarisation values, with a peak
of about 65% for Er and a clear even-odd disparity, where even values of Z bring
along an additional 10–20% of polarisation. While the two techniques involved
are probing with rather different energy resolution, they are both predominantly
14 Magnetism of the Elements 685

Fig. 16 The polarisation of the heavy rare earths. (TM data form Ref. [11], PE data from Ref. [20]
FE data from Ref. [17])

sensitive to the polarisation of the electrons at EF . The odd-even disbalance cannot


be related to the involvement of superconducting Cooper pairs in PCAR, as there
is no such a dependence in the FE experiments. The occupancy and the Fermi
pairing correlations of the 4f states should be therefore the culprit. Detailed, fully
relativistic band structure calculations may be necessary, in order to explain this
characteristic behaviour.

Oxygen

Molecular Magnetism

Molecule-Based Magnets vs. Single-Molecule Magnets


The majority of molecular magnets under active exploration contain individual
spins, often carried by Mn, Fe, Ni and V, which are only weekly exchange coupled
within a molecular framework. These are not typically ordering magnetically until
close to the absolute zero and as of the low moment density and their powder
or solution presentation are not useful as conventional magnetic materials. The
686 P. Stamenov

ability to associate individual quantum spins with a molecule, however, is of interest


to the fields of quantum information storage and quantum computing, so are the
phenomena of quantum tunnelling of the magnetisation and electron spin resonance
manipulation of individual molecules. Below certain blocking temperature, the
single-molecule magnets would exhibit magnetic hysteresis without the existence
of long-range magnetic order, of purely molecular, quantum-mechanical origin.
The first reports of single-molecule systems originate in the early 1990s of
the last century and are centred around manganese oxide compounds such as
the Mn12 O12 (OAc)16 (H2 O)4 , better known as Mn12 [22]. Recent efforts in the
field have pushed the blocking temperatures to above 65 K, using dysprosium
compounds, such as [Dy(CpiPr5 )(CpMe5 )][B(C6 F5 )4 ], still somewhat short of the
boiling point of liquid nitrogen.
Molecule-Based Moments
One interesting concept to explore is the level of localisation of the magnetic order in
systems which exhibit uncompensated spins. It is possible to contemplate magnetic
materials, where the structural building blocks are molecular in nature. These can
be comprised of organic molecules and metal-ion coordination compounds where
the unpaired electron spins may reside on p, d and f orbitals, which belong to an
entire molecule rather than an atom. These are able to display bulk ferromagnetic
and ferrimagnetic behaviour and have well-defined, be they often rather low, critical
temperatures (Neel or Curie points). This behaviour is in sharp contrast with the one
of single-molecule magnets, which are inevitably superparamagnetic and may only
display a blocking temperature. Probably the first historical example of a molecule-
based magnet (in this case S = 3/2) is the compound diethyldithiocarbamate-Fe(III)
chloride, synthesised and extensively characterised by Wickman and co-workers in
the late 1960s of the twentieth century [23].
The Magnetic Phase Diagram of Oxygen
Oxygen is in the rather special position to be the only element to be able to
form ordered antiferromagnetic crystals, which are Van der Waals insulators. The
antiferromagnetic interactions in this molecular system are understandably closely
dependent on the crystal structure. There are three main crystallographic phases
below the melting point of Tm = 54.4 K. The lowest temperature α − O2 phase
is monoclinic (C2/m), with all the O–O bonds order perpendicular to the basal
a − b plane. This gives way at a temperature of Tαβ = 23.8 K, to a rhombohedral
(R3̄m) lattice with a short-range magnetic order β − O2 . A further transition at a
Tβγ = 43.8 K leads to a higher symmetry cubic structure (Pm3n), where the O2
molecules are orientationally disordered, almost completely isotropic. The strong
interplay between crystal structure and magnetic order manifests itself as both
magnetostriction and shift of the transition temperatures in the presence of high
magnetic field. Shifts of the order of 1 K per 10 T are readily accounted for by
variations of the Landau theory [24].
An excellent review summarising the experimental efforts, from the early pro-
duction of liquid and solid oxygen to the complete characterisation of its structural
14 Magnetism of the Elements 687

and magnetic properties, spanning the best part of last century, is found in [25].
The susceptibility of α − O2 is anisotropic, with an easy axis along the primary
twofold direction of the monoclinic structure. Discrepancies between the sublattice
and intersublattice exchange constants extracted from the parallel and perpendicular
susceptibilities, from the antiferromagnetic resonance frequencies and from heat
capacity measurements suggest that the appropriate model requires multiple non-
collinear magnetic sublattices to account for all observations. The susceptibility of
β − O2 is essentially anisotropic with strong temperature dependence, due primarily
to the lattice expansion and a corresponding modulation of the exchange constants.
The paramagnetic susceptibility of γ −O2 does not follow the Curie-Weiss law, with
a substantially weaker temperature dependence than expected, due to a combination
of the existence of short-range correlations and the partial obstruction of free
rotation for the molecular units [25].
The relatively high susceptibility of liquid and solid oxygen is often an item
of practical experimental concern in high sensitivity magnetisation measurements
at cryogenic temperatures, as air can be easily admitted through various seals
in the sample space and condense in the vicinity of the tested sample [26]. In
order to measure the susceptibility of oxygen specimen with a sufficient accuracy,
early experiments have been typically conducted by absorption onto Vycor or
Graphite or on crystals prepared by directional solidification [27]. While thin partial
monolayer coverages exhibit transitions which are rather blurred, it takes only
four to five monolayers to establish the ordering at close to bulk parameters, with
sharp and well-defined critical temperatures, close to the ones shown in Fig. 17.
The magnetism here is strictly of molecular origin with the ground state 3 Σg−
exhibiting a spin of S = 1 and an orbital angular momentum of L = 0. The
next two available molecular states of spin S = 0 are sitting 0.98 eV (1 Δg ) and
1.63 eV (1 Σg+ ) above the ground state. This substantial spacing ensures negligible
population at temperatures where oxygen is solid [28]. The solid oxygen system
can be represented as antiferromagnetically interacting molecular spins, with rather
substantial anisotropy of K ≈ 16 J and therefore a gapped harmonic spin-wave
excitation spectrum Eg /kB = 80.5 K, as clarified by a set of muon spin rotation
(μSR) experiments [28].
More recent works focus on the completion of the magnetic and structural phase
diagram in the higher pressure-temperature region, with new phases, such as δ
and , requiring tighter control of experimental conditions in order to clarify how
antiferromagnetic coupling between nearest-neighbour O2 molecules results in the
formation of larger O4 units in a singlet state and a second-order phase transition to
an orthorhombic structure [29, 30]. The nature of the high-pressure phase (seen
in Fig. 18) is still currently under debate. The problem for the relative stability of
the couple of possible antiferromagnetic arrangements in four-molecule (tetramer
units, which are essentially flat and have opposing oxygen molecular moments
either sharing diagonal or a side in the unit rhombus) units versus the non-magnetic
solution is approached by a combination of high-accuracy DFT electronic structure
calculations and experimental observations of the evolution of lattice parameters
with pressure [31].
688 P. Stamenov

Fig. 17 The low-frequency susceptibility of solid oxygen. This has been measured at a frequency
of 155 Hz, with an excitation field of 0.19 mT, for a bulk specimen of 250.1 mg, below the boiling
point at ambient pressure Tb = 90.2 K. (After: [27])

Fig. 18 The high-pressure phase diagram of oxygen. The γ phase is paramagnetic. (After: [30])
14 Magnetism of the Elements 689

Other Examples of Magnetic Order in the p- and d-Shell Elements

Light p-Shell Elements: Carbon


A vast amount of publications exists with both theoretical and experimental claims
over the magnetism of p-shell elements, carbon magnetism , in particular [32]. Here
the focus is on magnetism in systems, which are, at least nominally, pure. In most
cases the claims are made based on one or more forms of magnetometry: bulk, most
often SQUID-based inductive magnetometry, scanning probe techniques, such as
magnetic force microscopy (MFM) and far less commonly on element-specific and
site-specific techniques. Quite possibly all allotropic forms of carbon have been,
at some point, alleged to be magnetic. For example, the famous C6 0 fullerene is
becoming weakly magnetic, when a single carbon atom in the buckyball structure
is replaced with nitrogen, but with a rather low and unattractive from applications
point of view curie point of about 15 K [33]. Braver claims have been made on
the ferromagnetism of disordered carbon nanoparticles and mixtures of nanorods,
nanotubes and nanospheres, produced by pulsed arc discharge, persisting well above
300 K [34]. Substantially more conservative Curie temperatures (of up to 90 K) have
been in carbon nano-foam produced by high-power laser oblation of glassy carbon
and attributed to the mixture of sp2 and sp3 hybridisation and the steric topological
protection offered by the convoluted sheets [35]. Proton irradiation of highly
oriented pyrolytic graphite (HOPG) has been argued to result in ferromagnetism
persistent above room temperature [36]. Probably the most substantiated evidence
for the magnetism of thin proton-irradiated carbon films comes from a combination
of MFM and XMCD spectromicroscopy [37]. Robust magnetic signals are found
after proton irradiation with a fluence of 50 nC/μm2 , while no contamination signal
of either Fe, Co or Ni could be resolved on the corresponding L3 edges, as seen
in Fig. 19. While this type of experimental evidence points towards the assertion
that the magnetism at least in some carbon forms is defect-related and not due
to extrinsic magnetic contamination, because of the vast variety of methods used
for sample preparation, the introduction of non-magnetic dopants, such as N, and
sample treatments, it is safe to assert that this debate is not yet closed.
Heavy p-Shell Elements
There are still to be verified theoretical predictions that a number of p-block
elements (especially heavier ones) absorbed as Bi-X dimers on surfaces can exhibit
giant magnetic anisotropy energies and correspondingly molecular magnetism [38].
Nanomagnetic units with anisotropy energies of the order of 200 meV are expected
on common and accessible substrates, such as MgO 100 .
Heavy d-Shell Elements
While bulk ruthenium is paramagnetic in its HCP cubic structure and has found
applications as a spacer in spin-valve structures, especially ones containing synthetic
antiferromagnets, it has been recently proposed and experimentally evidenced
that BCT ruthenium can be stabilised by the use of suitable seed and capping
690 P. Stamenov

Fig. 19 AFM/MFM images, together with XMCD spectromicroscopy images of the same proton-
irradiated area of a thin carbon film. (After: [37])

layers [39]. If independently confirmed, this would represent the first demonstration
of a single element 4d ferromagnet, with a transition point above room temperature,
appreciable saturation magnetisation of Ms 0.15 MA/m and useful anomalous Hall
resistance. There is even a theoretical possibility for the realisation of a substantial
perpendicular anisotropy, as high as a MJ/m3 , if an out-of-plane texturing is to
be achieved. A resurgence of the interest in similar systems, narrowly missing on
satisfying Stoner’s criterion, such as Pd and Pt, is to be expected.
Heavy f-Shell Elements
These are the four stable paramagnetic actinides: Th, Pa, U and Np. The f -electrons
are more delocalised than for the corresponding 4f elements, forming hybridised
energy bands and none of the four order magnetically, although there are some
magnetic alloys and compounds. As a result of the strong spin-orbit coupling, the
L − S scheme that gives Hund’s rule moments for the rare earths breaks down for
the actinides, where the J − J coupling scheme is more appropriate, as the electrons
become more localised with increasing atomic number.
14 Magnetism of the Elements 691

References
1. Coey, J.M.D.: Magnetism and Magnetic Materials. Cambridge University Press, Cambridge
(2010)
2. Coey, J.M.D.: Materials for spin electronics. In: LNP 569, pp. 277–297. Springer, Berlin/Hei-
delberg (2001)
3. Warnes, L., King, H.: The low temperature magnetic properties of austenitic Fe-Cr-Ni alloys
2. The prediction of Neel temperatures and maximum susceptibilities. Cryogenics 16(11), 659
(1976)
4. Hobbs, D.S.D., Hafner, J.: Understanding the complex metallic element Mn. I. Crystalline and
noncollinear magnetic structure of a-Mn. Phys. Rev. B 68, 014407 (2003)
5. Zabel, H.: Magnetism of chromium at surfaces, at interfaces and in thin films, J. Phys.:
Condens. Matter 11, 9303–9346 (1999)
6. Fawcett, E.: Spin-density-wave antiferromagnetism in chromium. Rev. Mod. Phys. 60, 209
(1988)
7. Fawcett, E., Alberts, H.L., Galkin, V.Y., Noakes, D.R., Yakhmi, J.V.: Spin-density-wave
antiferromagnetism in chromium alloys. Rev. Mod. Phys. 66, 25 (1994)
8. Jensen, J., Mackintosh, A.R.: Rare Earth Magnetism – Structures and Excitations. Clarendon
Press, Oxford (1991)
9. Houmann, J.G., Rainford, B.D., Jensen, J., Mackintosh, A.R.: Magnetic excitations in
praseodymium. Phys. Rev. B 20, 1105 (1979)
10. Plessis, P.D.V.D.: Magnetic anisotropy of some heavy rare-earth metals. Physica 41, 379 (1969)
11. Mazin, I.I.: How to define and calculate the degree of spin polarization in ferromagnets. Phys.
Rev. Lett. 83, 1427 (1998)
12. Stamenov, P.: Point contact andreev reflection from erbium: the role of external magnetic field
and the sign of the spin polarization. J. Appl. Phys. 111, 07C519 (2012)
13. Min, B.I., Jang, Y.R.: The effect of the spin-orbit interaction on the electronicstructure of
magnetic materials. J. Phys.: Condens. Matter 3, 5131 (1991)
14. Allenspach, R., Mauri, D., Taborelli, M., Landolt, M.: Spin-polarized Auger-electron spec-
troscopy. Phys. Rev. B 35, 4801 (1987)
15. Shirokovskii, V.P., Kirillova, M.M., Shilkova, N.A.: An optical absorption anomaly in iron.
Sov. Phys. JETP 55, 464 (1982)
16. Busch, G., Campagna, M., Siegmann, H.C.: Spin-polarized photoelectrons from Fe, Co, and
Ni. Phys. Rev. B 4, 746 (1971)
17. Chrobok, G., Hofmann, M., Regenfus, G., Sizmann, R.: Spin polarization of field-emitted
electrons from Fe, Co, Ni, and rare-earth metals. Phys. Rev. B 15, 429 (1976)
18. Valentine, J.M., Chien, C.L.: Determination of spin polarization of Gd and Dy by point-
contactAndreev reflection. J. Appl. Phys. 99, 08P902 (2006)
19. Stamenov, P., Coey, J.: Fermi level spin polarization of polycrystalline thulium by point contact
Andreev reflection spectroscopy. J. Appl. Phys. 109, 07C713 (2011)
20. Meservey, R., Paraskevopoulos, D., Tedrow, P.M.: Tunneling measurements of conduction-
electron-spin polarizationin heavy rare-earth metals. Phys. Rev. B 22, 1331 (1980)
21. Fecher, G., Morais, J., Liesegang, J., Braun, J., Cherepkov, N., Oelsner, A., Günther, M.,
Schicketanz, M., Schönhense, G.: Spin polarisation and dichroism in ARUPS from thin rare
earthfilms. J. Electr. Spec. 114, 1171 (2001)
22. Caneschi, A., Gatteschi, D., Sessoli, R.: Alternating current susceptibility, high field
magnetization, and millimeter band EPR evidence for a ground S = 10 State in
[Mn12012(CH3C00)16(H20)4].2CH3C00H.4H20. J. Am. Chem. Soc. 113, 5873–5874 (1991)
23. Wickman, H.H., Trozzolo, A.M., Williams, H.J., Hull, G.W., Merritt, F.R., Spin-3/2 iron
ferromagnet: its mossbauer and magnetic properties. Phys. Rev. 163, 526 (1967)
24. Gomonay, E.V., Loktev, V.M.: Shift of the basal planes as the order parameter of transitions
between the antiferromagnetic phases of solid oxygen. Low Temp. Phys. 31(8–9), 763 (2005)
25. DeFotis, G.C.: Mgnetism of solid oxygen. Phys. Rev. B 23(9), 4714 (1981)
692 P. Stamenov

26. Mpms application note 1014-210 oxygen contamination


27. Gregory, S.: Magnetic susceptibility of oxygen adsorbed on graphite. Phys. Rev. Lett. 40(11),
723 (1978)
28. Morris, G., Brewer, J., Dunsiger, S., Montour, M.: Antiferromagnetism in solid oxygen.
Hyperfine Interact. 104, 381–385 (1997)
29. Gorelli, F., Ulivi, L., Santoro, M., Bini, R.: Antiferromagnetic order in the d phase of solid
oxygen. Phys. Rev. B 62(6), R3604 (2000)
30. Gorelli, F.A., Santoro, M.: High-pressure antiferromagnetic phases of solid oxygen. J. Raman
Spectrosc. 34, 549–556 (2003)
31. Ramírez-Solís, A., Zicovich-Wilson, C.M., Hernández-Lamonedab, R., Ochoa-Callec, A.J.:
Antiferromagnetic vs. non-magnetic e phase of solid oxygen. Periodic density functional
theory studies using a localized atomic basis set and the role of exact exchange. Phys. Chem.
Chem. Phys. 19, 2826 (2017)
32. Volnianska, O., Boguslawski, P.: Magnetism of solids resulting from spinpolarization of p
orbitals. J. Phys.: Condens. Matter 22, 073202 (2010)
33. Simon, H.A.F., Náfrádi, B., Fehér, T., Forró, L., Fülöp, F., Jánossy, A., Korecz, L., Rock-
enbauer, A., Hauke, F., Hirsch, A.: Magnetic fullerenes inside single-wall carbon nanotubes.
Phys. Rev. Lett. 97, 136801 (2006)
34. Parkansky, N., Alterkop, B., Boxman, R.L., Leitus, G., Berkh, O., Barkay, Z., Rosenberg, Y.,
Eliaz, N.: Magnetic properties of carbon nano-particles produced by a pulsed arc submerged
in ethanol. Carbon 46, 215 (2008)
35. Rode, A.V., Gamaly, E.G., Christy, A.G., Gerald, J.G.F., Hyde, S.T., Elliman, R.G., Luther-
Davies, B., Veinger, A.I., Androulakis, J., Giapintzakis, J.: Unconventional magnetism in all-
carbon nanofoam. Phys. Rev. B 70, 054407 (2004)
36. Esquinazi, P., Spemann, D., Höhne, R., Setzer, A., Han, K.H., Butz, T.: Induced magnetic
ordering by proton irradiation in graphite. Phys. Rev. Lett. 91, 227201 (2003)
37. Ohldag, H., Tyliszczak, T., Höhne, R., Spemann, D., Esquinazi, P., Ungureanu, M., Butz, T.:
π -electron ferromagnetism in metal-free carbon probed by soft X-ray dichroism. Phys. Rev.
Lett. 98, 187204 (2007)
38. Pang, R., Deng, B., Shi, X., Zheng, X.: Giant magnetic anisotropy of heavy p-elements on high-
symmetrysubstrates: a new paradigm for supported nanostructures. New J. Phys. 20, 043056
(2018)
39. Quarterman, P., Sun, C., Garcia-Barriocanal, J., Mahendra, D.C., Lv, Y., Manipatruni, S.,
Nikonov, D.E., Young, I.A., Voyles, P.M., Wang, J.P.: Demonstration of Ru as the 4th
ferromagneticelement at room temperature. Nat. Commun. 9, 2058 (2018)

Dr. Stamenov received a BSc degree from the University of


Sofia in 2002, with research work on manganese perovskites. He
completed his PhD in 2007 (Prof. J. M. D. Coey as supervisor)
in Trinity College Dublin, focusing on metals, semimetals and
semiconductors for spin electronics applications. In 2010, he
became a Lecturer in Physics and Principal Investigator within
CRANN in the area of nanomagnetism.
Metallic Magnetic Materials
15
J. Ping Liu , Matthew Willard , Wei Tang, Ekkes Brück ,
Frank de Boer, Enke Liu , Jian Liu, Claudia Felser ,
Gerhard Fecher, Lukas Wollmann, Olivier Isnard, Emil Burzo,
Sam Liu, J. F. Herbst , Fengxia Hu, Yao Liu, Jirong Sun,
Baogen Shen, and Anne de Visser

Contents
Magnetic Metallic Glasses and Nanocrystalline Alloys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 695
Alnicos . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 698
Intermetallic Compounds of d-d and d-p Types . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 705
Magnetic Shape Memory Alloys and Compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 705
Magnetic Heusler Compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 717
Intermetallic Compounds of 3d-4f Type . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 727
Sm-Co Permanent Magnets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 757
Nd-Fe-B Permanent Magnets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 760
Magnetocaloric Intermetallics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 766
Heavy-Fermion Compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 771
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 778

Abstract

Metallic magnetic materials are based on the ferromagnetic metallic elements,


and their alloys or intermetallic compounds, though non-ferromagnetic elements,
may also play key roles in some of the ferromagnets. Metallic magnetic materials
encompass many of the materials of greatest technological importance, including

J. P. Liu
University of Texas at Arlington, Arlington, TX, USA
e-mail: pliu@uta.edu
M. Willard
Materials Science and Engineering, Case Western Reserve University, Cleveland, OH, USA
e-mail: maw169@case.edu
W. Tang
Materials Science and Engineering, Ames Laboratory, Ames, IA, USA
e-mail: weitang@ameslab.gov

© This is a U.S. Government work and not under copyright protection in the U.S.; 693
foreign copyright protection may apply 2021
J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_16
694 J. P. Liu et al.

soft magnets and permanent magnets for energy conversion applications, as


well as materials for functions such as magnetostriction, shape memory, and
magnetorefrigeration.
This chapter, written by a team of authors, is organized as follows: we
begin with Sect. “Magnetic Metallic Glasses and Nanocrystalline Alloys” on
amorphous iron- and iron–cobalt-based soft ferromagnetic materials and then,
in Sect. “Alnicos”, Alnicos, an important class of hard magnetic materials
(permanent magnets) based on shape anisotropy of aligned nanoscale iron–cobalt
needles embedded in a nonmagnetic matrix. These two sections illustrate how
magnetic performance is related to specific phase structures and morphology of
the ferromagnetic elements or simple alloys. However, most advanced magnetic
materials are based on intermetallic compounds with complex crystalline and
electronic structures where electronic interactions often govern the magnetic
properties. Section “Intermetallic Compounds of d-d and d-p Types” intro-
duces the d-d and d-p intermetallic compounds. Then, Sect. “Magnetic Shape

E. Brück
Delft University of Technology, Delft, The Netherlands
e-mail: e.h.bruck@tudelft.nl
F. de Boer
University of Amsterdam, Amsterdam, The Netherlands
E. Liu · F. Hu · Y. Liu · J. Sun · B. Shen
Institute of Physics, Chinese Academy of Sciences, Beijing, China
e-mail: ekliu@iphy.ac.cn; fxhu@iphy.ac.cn; shenbg@iphy.ac.cn; jrsun@g203.iphy.ac.cn
J. Liu
Ningbo Institute of Materials Technology and Engineering, Chinese Academy of Sciences,
Ningbo, China
e-mail: liujian@nimte.ac.cn
C. Felser · G. Fecher · L. Wollmann
Max-Planck-Institute für Chemische Physik fester Stoffe, Dresden, Germany
e-mail: Claudia.Felser@cpfs.mpg.de; Gerhard.Fecher@cpfs.mpg.de
O. Isnard
Institute Néel and Université Grenoble Alpes, Grenoble, France
e-mail: olivier.isnard@grenoble.cnrs.fr
E. Burzo
Babes-Bolyai University, Romania, Cluj-Napoca, Romania
e-mail: emil.burzo@phys.ubbcluj.ro
S. Liu
University of Dayton, Dayton, OH, USA
J. F. Herbst
Research & Development, General Motors R&D Center, Warren, MI, USA
A. de Visser
Van der Waals-Zeeman Institute, University of Amsterdam, Amsterdam, The Netherlands
e-mail: A.deVisser@uva.nl
15 Metallic Magnetic Materials 695

Memory Alloys and Compounds” discusses magnetic shape memory materials


that are dependent on d-electrons, and Sect. “Magnetic Heusler Compounds”
summarizes recent progress in the magnetic intermetallic Heusler compounds.
The 4f electrons are so important in forming magnetocrystalline anisotropy
that is the key property for many hard magnetic materials. A comprehensive
review on the 3d-4f intermetallic compounds is provided in Sect. “Intermetallic
Compounds of 3d-4f Type”, followed by Sects. “Sm-Co Permanent Magnets”
and “Nd-Fe-B Permanent Magnets” on specific discussions of two major types
of rare-earth permanent magnets based on Sm-Co and Nd-Fe-B intermetallics.
Section “Magnetocaloric Intermetallics” reviews magnetocaloric compounds,
and Sect. “Heavy-Fermion Compounds” is devoted to heavy-fermion compounds
where 5f electron interactions often define the magnetic and other physical
properties. The authors of each section are the leading researchers in the specific
topic. The chapter is organized and edited by J. Ping Liu, with editorial assistance
from Jeotikanta Mohapatra and Meiying Xing of the University of Texas at
Arlington.

Magnetic Metallic Glasses and Nanocrystalline Alloys

The irregular atomic arrangement in amorphous ferromagnetic metals results in


reduced anisotropy, coercivity (Hc ), Curie temperatures (TC ), and magnetizations
when compared to the crystalline counterparts. Nanocrystalline soft magnetic
materials have been developed with higher magnetization and, comparably, small
coercivities. Together, the amorphous and nanocrystalline materials exhibit desir-
able characteristics for a wide range of applications with switching in the sub-MHz
frequency range [1]. At these frequencies, eddy currents add to the overall losses;
however, they can be mitigated to some extent by selection of materials with
large resistivity (ρ). The total losses due to the switching of the magnetic material
are called the core losses (Pcm ) and are graphically visualized as the area within
the B-H hysteresis loop. In this section, the characteristics unique to amorphous
and nanocrystalline soft magnets, the nonequilibrium methods required for their
preparation, and their outstanding magnetic properties will be presented.
The amorphous alloys or metallic glasses are characterized structurally by the
absence of long-range periodic atomic order. Samples with thicknesses exceeding
approximately 1 mm are referred to as bulk amorphous materials with no long-range
periodic order or bulk metallic glasses. Crystallization is avoided by sufficiently
rapid cooling from the melt. For conventional amorphous alloys, the cooling rate
is >105 K/s, while for bulk amorphous materials, the critical cooling rate may
be as low as 10 K/s. The elements used to create amorphous, bulk amorphous,
and nanocrystalline soft magnetic alloys have many commonalities. As shown in
Fig. 1, all three groups are made up of a majority of 3d magnetic transition metals,
especially Fe, Co, and Ni, which give the desired magnetization.
696 J. P. Liu et al.

Fig. 1 Elemental makeup of typical nanocomposite soft magnetic alloys with four major compo-
nents: magnetic transition metals (MTM), early transition metals (ETM), metalloid/post-transition
metals (PTM), and late transition metals (LTM)

Glass-forming ability is imparted by the metalloid and post-transition metal


elements in most amorphous and bulk amorphous alloys. This is largely due to the
partitionless transformation to the amorphous phase that can occur directly from the
melt at eutectic compositions near pure MTM. Early transition elements are used
in some amorphous and bulk amorphous and in all nanocrystalline soft magnetic
alloys. The ETMs are necessary as a deterrent to grain coarsening in nanocrystalline
alloys, and they are usually accompanied by late transition metal alloying elements,
which provide nucleation sites during the early stages of crystallization [2, 3].
The conventional melt spinning technique uses an induction coil to melt ingots of
a desired composition and then expels the melt through an orifice in the crucible onto
a rapidly rotating wheel, giving effective quench rates of up to 106 K/s. Using this
conventional technique, alloys with compositions near deep eutectics are routinely
formed into amorphous ribbons 20 to 50 microns in thickness, a few millimeters
to many centimeters in width, and meters in length. Bulk amorphous alloys can be
formed at lower cooling rates by casting, sometimes with a flux to prevent oxidation
and improve the maximum casting diameter.
Nanocrystalline alloys can be produced using an amorphous alloy as precursor,
but most studies have been carried out on melt-spun ribbons with post-quench
annealing to optimize alloy performance. The final microstructure consists of
nanocrystalline grains surrounded by a thin amorphous matrix phase (1–2 nm in
width). This gives the optimal microstructure with 70–80 vol% crystallinity. More
details about the processing of amorphous, bulk amorphous, and nanocrystalline
alloys can be found in reviews [4–6].
The MTM composition has a major influence on the magnetization, magne-
tostriction, magnetocrystalline anisotropy, and Curie temperature. Adjusting the
ratios of Co, Ni, and Fe influences the crystallization behavior and magnetic
performance. While Fe-based alloys have the great advantages of low cost and
strong performance in terms of magnetization and low losses, improvements can
be made by adjusting the composition to increase the operating temperature and
saturation magnetization. The Curie temperature of the amorphous phase (TCam ) is
an upper bound on the operating temperature of all three classes of alloys.
In amorphous alloys, the lack of long-range periodic order tends to minimize
the local magnetocrystalline anisotropy, leaving magnetoelastic anisotropy and
long-range dipolar anisotropy to dominate. In nanocrystalline alloys, exchange
averaging helps to reduce the effective magnetocrystalline anisotropy, leaving the
magnetostriction as the major influence on the core losses. Residual stress in
15 Metallic Magnetic Materials 697

Fig. 2 Magnetocrystalline anisotropy (K1 ) energy density for BCC Fe single crystal (dark blue)
as a function of angle from the [001] direction when viewed from the [100] projection and the
effective magnetocrystalline anisotropy (light green) for nanostructured BCC Fe (where N grains
are exchange coupled in a volumetric exchange length (Lex 3 ))

the alloy couples magnetoelastically with the magnetization resulting in increased


losses. This can be avoided by choosing compositions where the magnetostrictive
coefficient (λ) is near zero.
The remarkable reduction in coercivity when grain sizes are reduced to the
nanocrystalline regime was first described by Herzer [7]. Based on the idea that
randomly oriented grains are exchange coupled through the residual amorphous
matrix, a random anisotropy model shows that the coercivity is proportional to
the grain size to the sixth power. This is possible when the exchange correlation
length (Lex ) is larger than the structural correlation length (grain size, D), an effect
referred to as exchange softening which results from exchange averaged anisotropy.
This is illustrated in Fig. 2, where the integrated anisotropy energy is unchanged
by the averaging but the fluctuations (which are more important to soft magnetic
properties) are significantly reduced.
The exchange length indicates the minimum length scale over which the atomic
moments must remain aligned due to exchange forces. The magnitude √ of this
fundamental magnetic material parameter can be found by Lex = A/K1 , where
A is the exchange stiffness and K1 is the magnetocrystalline anisotropy constant.
A value of 35 nm was calculated for Fe-Si-based nanocrystalline alloys. Since
D ∼10 nm is much smaller than Lex , the magnetic moments in each individual
grain cannot relax into the local easy direction dictated by the grain orientation.
This results in averaging of the local magnetocrystalline anisotropy over the
exchange correlation volume (Lex 3 ). Using the random anisotropy model, the
effective magnetocrystalline anisotropy,
√ K1 , representing the material response
can be determined as K = K1 / N, with N being the number of grains within
698 J. P. Liu et al.

the exchange correlation length. The value of N in a cubic volume with sides
Lex can be estimated by the relation N = (Lex /D)3 . Using these equations and
the definition of Lex , the effective anisotropy can be determined in terms of the
crystalline material parameters, K1 , A, and D: K = K14 · D 6 /A3 . The exchange
softening described by this model breaks down as the grains become decoupled,
especially when the Curie temperature of the amorphous intergranular phase is
exceeded. This effective anisotropy is proportional to the coercivity and inversely
proportional to the permeability of nanocrystalline alloys.
Due to the small value of the averaged anisotropy, amorphous and bulk amor-
phous alloys have relatively small coercivity (again proportional to K1 ) [8–11]. In
these materials, the magnetoelastic anisotropy dominates the coercivity. An unfortu-
nate correlation between the square of magnetization and the magnetostrictive coef-
ficient is found for amorphous alloys. This effect is absent in nanocrystalline alloys
due to the two-phase microstructure which can be adapted to provide near-zero
values of magnetostrictive coefficient together with moderate magnetization [12].
There are a number of advantages that amorphous and nanocrystalline soft
magnetic alloys have over conventional large-grained materials, as summarized
in Fig. 3 and Table 1. Among the advantages, an important one is their better
performance at switching frequencies above 1 kHz. This is due in part to (a) the
smaller interaction of domain walls with the nanocrystalline grains and (b) their
higher resistivity. The hysteresis loss is present at all frequencies and proportional
to frequency. Controlling the microstructure and composition of the alloy can lead to
reduction of hysteretic losses. As the switching frequency is increased, eddy currents
in the alloy enhance the core losses. Reducing lamination thickness and increasing
the resistivity of the alloy help to limit the eddy currents. At high frequencies
(especially above 10 kHz), dynamic eddy current losses dominate. These eddy
currents are localized at the domain walls and can be minimized by refinement of
the domain structure in the alloys.

Alnicos

In the first section, iron–cobalt-based alloys featured as soft magnetic materials.


Here, iron–cobalt-based nanostructures feature as hard magnets. Alnico magnets
have been developed, based on ferromagnetic Alnico alloys discovered by T.
Mishima in 1932 [17]. They are composed of Fe, Co, Ni, and Al with additions of
Cu, Ti, and Nb. It is generally recognized that high coercivity Hc in Alnico magnets
depends on the shape anisotropy of the ferromagnetic Fe-Co precipitates (α 1 -
phase) finely dispersed in a nonmagnetic Ni-Al matrix (α 2 -phase). This dispersion
is caused by spinodal decomposition (SD) through an appropriate heat treatment
[18–27].
The most widely used Alnico alloys are Alnico 5–7, 8, and 9 (Table 2).
Micrographies showing typical microstructure of a commercial Alnico 9 magnet
are shown in Fig. 4 where the high-aspect-ratio, aligned iron–cobalt needles can
be seen [27]. The longitudinal microstructure (a) exhibits highly elongated α 1
rods (>400 nm) which produce a strong shape anisotropy. The transverse (b)
consists of the Fe-Co-rich α 1 -phases (brighter mosaic tile), Ni-Al-rich α 2 -phases
15 Metallic Magnetic Materials 699

Fig. 3 The magnetic permeability at 1 kHz plotted against the saturation induction (μ0 Ms ) for
soft magnetic materials [8–11]

(darker matrix), and Cu-rich phases (white spots located between the corners of
α 1 -phase). The transverse microstructure of Alnico 8 is similar to that of Alnico
9, but the longitudinal α 1 -phase is elongated in two orthogonal <100> by the
applied field [27]. The magnets can be grouped into two main categories: isotropic
and anisotropic, distinguished by the different alloy crystallization processes and
thermomagnetic treatments.
700

Table 1 Magnetic properties of commonly used soft magnetic materials [13–16]. Metglas is amorphous, Senntix is bulk amorphous, and Finemet is
nanocrystalline. The saturation magnetization for alloys marked with “†” exceeds 1 T. Core loss (Pcm ) is reported for 60 Hz sinusoidal switching at 1T
maximum induction amplitude. Js is the magnetic polarization at saturation; μmax is the maximum permeability; Tx1 is the primary crystallization temperature
 (10−8 
Material Js (T) μmax (103 ) Hc (A/m) m) TC (K) Pcm (W/kg) λs (ppm) Tx1 (K)
Fe49 Co49 V2 2.4 5–50 16–398 27 1203 1.10 60
Soft Fe 2.16 10–50 4–80 10 1044 20 −2
Fe 3.2%Si 2.0 0.5–5 6 48 1018 0.84 7
Fe67 Co18 B14 Si1 1.8 50 3.5 123 658 0.66 35 703
Metglas 2605CO
Fe81 B14 Si3 C2 1.56 600 0.1 130 668 0.08 27 783
Metglas 2605SA1
Ni50 Fe50 1.4–1.6 70 4–20 40–50 753 0.33 18
Permalloy
Fe77 P7 B13 Nb2 Cr 1.31 2.5 556 0.65 W/m3 795
Senntix
Fe73.5 Si13.5 Nb3 B9 Cu 1.23–1.35 80 0.6–2.5 110 843 0.01 0–2 780
Finemet
Fe40 Ni38 Mo4 B18 0.88 800 0.4 160 626 † 12 683
Metglas 2826MB
Co70 Fe5 Ni2 B3 Si15 0.77 600 0.8 136 638 † 0 793
Metglas 2705M
Ni78 Fe17 Mo5 0.65–0.82 100–800 0.25–0.64 60 673 † 2–3
Supermalloy
J. P. Liu et al.
Table 2 Nominal composition and magnetic properties of cast Alnico 8 and 9 magnets [28, 29]
No. Characteristics Type Composition (wt.%, balanced Fe) Magnetic properties
Co Ni Al Cu Others Br(T) Hc (kA/m) (BH)max (kJ/m3 )
1 A 5–7 24.3 14.0 8.2 2.3 1.0 Nb 1.35 58.9 60.0
15 Metallic Magnetic Materials

2 B 8 40.1 13.0 7.1 3.0 6.5 Ti 0.74 151.2 41.0


3 A 9 35.4 12.1 7.0 3.2 5.0 Ti, 0.5 Nb 1.12 109.5 81.0
4 A 9 36 13.5 7.2 3.0 6.05 Ti 1.08 125 90.0
5 A 9h 40 14 8.4 3.0 8.2 Ti 0.86 160.8 75.6
6 A 9 34 – – – – 1.14 117.0 93.5
7 A 9 36 – – – – 1.13 126.0 104.3
8 A 9 38 – – – – 1.08 132.0 97.9
9 A 9 40 – – – – 1.04 144.0 93.4
10 B 8 38 – – – 6.5 Ti 0.88 147.0 47.8
11 B 8 38 – – – 7.5 Ti 0.80 152.0 46.2
12 B 8 38 – – – 7.5 Ti, 0.5 Nb 0.77 183.0 44.6
13 B 8 38 8.0, 2.0 Nb 0.61 209.3 34.0
Note: “A” field-treated, oriented grain; “B” field-treated, random grain; “–” unlisted or unknown
701
702 J. P. Liu et al.

Fig. 4 Microstructure images of a commercial Alnico 9 magnet (a) longitudinal, (b) transverse
[27]. (With kind permission from Elsevier)

Much work has been done on the composition, targeted at enhancing magnetic
properties, especially coercivity. The early studies on the relationship of columnar
crystallization conditions to compositions obtained a Br of 0.89–1.0 T, Hc of 141.7–
167.2 kA/m, and (BH)max of 87.6–93.1 kJ/m3 in the grain-oriented alloy containing
about 39 wt.% Co and 7–8 wt.% Ti [30]. By varying the Co and Ti content and
optimizing the microstructure, the magnetic properties of Alnico 9 were improved
as shown by No. 4–5 [29] in Table 2. Alnico 9 magnets are obtained by grain
orientation of Alnico 8 alloys with higher Co content [28]. By increasing the Co
content from 34 to 40 wt.%, as shown by No. 6–9 [28] in Table 2, Br decreases,
while Hc increases. A record (BH)max for Alnico of 104.3 kJ/m3 was achieved in
the magnet with 36 wt.% Co.
Several studies have confirmed that the addition of Ti effectively improves the
coercivity by increasing the local shape anisotropy constant [31], altering the lattice
mismatch between α 1 - and α 2 -phases [32], or narrowing the miscibility gap between
α 1 - and α 2 -phases [33]. Simultaneously increasing Co from 24 to 39 wt.% and Ti
from 0 to 7.5 wt.% resulted in an increase of Hc from 56.6 to 166.4 kA/m [31].
The addition of Nb showed a similar effect to that seen in Ti [34]. The addition
of 1.5 wt% Nb increased Hc by ∼ 24 kA/m. Moreover, Nb was found to help the
growth of columnar grains and suppress the precipitation of undesirable γ phase
[35]. Co-addition of Ti and Nb can significantly improve the coercivity, as shown
by No. 10–13 in Table 2, with a record Hc of 209.3 kA/m. Although Ti can improve
the coercivity, it makes the columnar crystallization of Alnico 9 difficult due to
fine-grained size. Therefore, a small amount of sulfur, selenium, or phosphorus
can be added to increase grain size and improve columnar crystal growth [36]. Cu
addition is necessary to separate the well-formed mosaic Fe-Co structure by forming
a Cu-rich phase [37], as shown in Fig. 4b. The Cu-rich phase forms as rods of 2
15 Metallic Magnetic Materials 703

Table 3 Magnetic properties of HP, HIP, and CMS samples [41]


Method Process Br (T) Hc (kA/m) (BH)max (kJ/m3 ) Grain size(μm)
HP 810 ◦ C, 10 min 0.91 142.9 45.4 90
HIP 1250 ◦ C, 4 h 0.94 146.9 47.0 50
CMS-1 1250 ◦ C, 4 h 0.88 134.1 39.8 30
CMS-2 1250 ◦ C, 8 h 1.01 134.4 51.7 330

nm diameter between α 1 -phase regions [27]. Cu was also thought to increase the
diffusion rate of atoms in the α 2 -phase [33].
Alnico magnets are manufactured using one of two processes: casting by
conventional foundry methods and powder metallurgical processing. The grain
orientation of cast Alnico 8 is random, whereas Alnico 9 with a perfect columnar
grain orientation is obtained by directional solidification. All the Alnico magnets
can be made from a powdered mixture of chemical elements or pre-alloyed powder
between 1250 ◦ C and 1350 ◦ C, but the grain orientation of magnets is random
[38, 39]. The magnetic alignment of cast and sintered magnets is achieved by
thermomagnetic treatment.
Gas atomization (GA) is an efficient technique to make Alnico powder [39]. The
powder is oval in shape and has an average size of 119 m. [40]. The powder can
be consolidated by techniques including hot pressing (HP), hot isostatic pressing
(HIP), and compression molding with subsequent sintering (CMS) [39, 41]. Table 3
shows the magnetic properties and grain size of HP, HIP, and CMS samples made
from the GA powder [41]. These magnets showed comparable magnetic properties
to cast magnets with the same composition. Injection-molded magnets are prepared
by mixing as-milled Alnico 8 powder with a binder [38]. The technique may allow
sintered magnets to be made with complex shapes.
A typical heat treatment process is shown in Fig. 5, where the alloys are first
solutionized (homogenized) at 1250 ◦ C for at least 30 min and then quenched to
obtain a single α-phase. Subsequently, the alloys are heat treated in a magnetic field
of 1.0–1.5 T, between 810 ◦ C and 840 ◦ C for 10–20 min to initiate the spinodal
decomposition. As a result, the α-phase separates into α 1 - and α 2 -phases, leading
to an initial coercivity. The applied magnetic field makes the Fe-Co particles to
preferentially grow in the field direction. A tempering treatment is performed at
lower temperatures for 15–30 h in two or more steps to promote chemical diffusion
between the α 1 - and α 2 -phases. These processes further increase the magnetic
anisotropy of the α 1 -phase, leading to enhanced coercivity.
The effects of thermomagnetic treatment at different conditions on microstruc-
ture and magnetic properties have been extensively investigated [20, 24, 26, 34,
37, 42, 43]. Iwama’s studies on thermomagnetic processing of Alnico magnets
established a linear relationship between a square of the wavenumber of the
modulated structure and the annealing temperature [44]. The relationship was
consistent with Cahn’s theoretical predictions on spinodal decomposition [18].
More recent studies confirm that achieving the highest Hc is closely related to
704 J. P. Liu et al.

Fig. 5 Schematic of the typical heat treatment for Alnico alloys

3.0
Fe – V
ATOMIC MOMENT IN BOHR MAGNETONS

Fe – Cr
2.5 Fe – Ni
Fe – Co
Ni – Co
2.0 Ni – Cu
Ni – Zn
Ni – V
1.5
Ni – Cr
Ni – Mn
1.0 Co – Cr
Co – Mn
PURE
0.5 METALS

0
Cr Mn Fe Co Ni Cu
6 7 8 9 10 11
ELECTRON CONCENTRATION, C

Fig. 6 Average atomic moments of binary alloys of the iron-group elements. (Reprinted figure
with permission from [47] by the American Physical Society)

optimizing the microstructures [26]. Chu’s studies explained that there exist two
splittings of Fe-Co particles during magnetic annealing which leads to a decrease of
mean diameter of Fe-Co particles and thus increases coercivity [42]. High magnetic
fields (up to 10 T) strongly influence the modulated (or mosaic) microstructure [43]
but did not improve Hc .
15 Metallic Magnetic Materials 705

Intermetallic Compounds of d-d and d-p Types

Binary intermetallic compounds show a great variety of magnetic properties.


Itinerant magnetism with non-integer magnetic moments was first observed and
explained in binary alloys of late 3d metals by Slater and Pauling [45, 46]. A few
years later, Bozorth demonstrated that the band picture becomes much more diverse,
as depicted in Fig. 6, if one does not restrict the alloying elements to direct neighbors
in the periodic table [47].
Magnetic ordering is not restricted to alloys of the magnetic 3d elements. This
section is the result of a thorough literature search for magnetically ordered binary
intermetallic d-d- and d-p-type compounds and alloys. The available compounds
with the type of magnetic order, the ordering temperature, and the value of the
magnetic moment are listed in Table 4. In the case of polymorphs, we include the
magnetic characteristics of all known crystal structures. We restrict ourselves to
alloys with metallic character and do not include alloys with rare-earth or actinide
elements.

Magnetic Shape Memory Alloys and Compounds

Magnetic shape memory (MSM) alloys and compounds are ferromagnets that
exhibit large strains under an applied magnetic field due to martensitic phase
transformation [173–175]. Currently, the materials with the best MSM properties
are the near-stoichiometric Ni2 MnGa compounds (a Heusler alloy) with five- and
seven-layer stacked martensite structures that show magnetic-field-induced strains
(MFIS) of 6.0% and 9.5% in a field less than 800 kA/m, respectively. This exceeds
the normal magnetostriction based on magnetoelasticity by orders of magnitude.
The origin of the MFIS in MSM alloys is due to magnetically induced twin boundary
motion in the martensitic phase. The MSM operation temperature is between
150 and 330 K depending on composition. These characteristics render the MSM
alloys promising functional materials for the development of magneto-mechanically
controlled actuators or dampers. The composition, crystal structure and struc-
tural/magnetic transition temperatures, and field-induced strain in single-crystalline
Ni-Mn-Ga alloys are given in Table 5. The characteristics and performance of Ni-
Mn-Ga alloys in other forms of polycrystals, composites, and thin films are collected
in Table 6.
In the second part of this section, the multifunctional MSM systems such as Ni-
Mn-(In, Sn, Sb) and Ni-Co-Mn-Ti Heusler alloys and hexagonal MM’X compounds
are addressed. Magnetic-field-induced reversible martensitic transformation has
been achieved in newly developed alloys, where a metamagnetic transition occurs
from weak magnetic martensite (austenite) to ferromagnetic austenite (martensite)
by applying a magnetic field. These alloys have the following advantages: (i) shape
recovery can be obtained in polycrystals, (ii) high blocking stress can be obtained,
and (iii) multifunctionality with large magnetocaloric effect and magnetoresistance
706

Table 4 Structure and magnetic properties of d-d and d-p types of intermetallic compounds (F stands for ferromagnetic, FI for ferrimagnetic, AF for
antiferromagnetic, SDW for spin density wave, and H for helimagnetic order. We include P for paramagnets if peculiar observations, e.g., metamagnetism,
are reported. The values given for the magnetic moment are per atom if not stated differently. Compositions can be either stated as weight % or atom a/o. RT
stands for room temperature, MCE for magnetocaloric effect, TC for Curie temperature, TN for Néel temperature sol. soln for solid solution and stoichiom fo
stoichiometry). Compounds are listed in order of 3d element
System Compound Type of order TC /TN (K) μ(μB ) Comment Ref.
Sc-Fe ScFe2 F 542 1.37 C14 Laves phase [48]
F >300 2.1 C36 Laves phase [49]
Sc-In Sc3 In F 16 0.06 [50]
Sc-Co ScCo2 P Metamagnet [51]
Ti-Be TiBe2 AF 10 1.64 C15 Laves phase [52]
Ti-Fe TiFe2 AF 0–65 C14; 67a/o Fe-64.5a/o Fe [53]
F 424–323 2.1–1.1 C14; 71.4a/o Fe-69.5a/o Fe [54]
Ti-Co TiCo3 F <80 Excess Co [55]
TiCo2 F 44 0.12 C36 [56]
AF 43 C15 [56]
Ti-Ni (Ni,Ti) F 627–0 0.6–0 15% Ti results in vanishing moment [57]
V-Fe σ -phase F 210–100 <0.2 D8b weak itiner. F 40–54a/o [V] [58]
α (V, Fe) F 0.6–1.3 B2 superstructure [59]
V-Co (V, α-Co) F Steep decrease of TC [57]
<RT @ 25%[V]
V-Ni (V, Ni) F Steep decrease of TC [57]
<RT @ 6%[V]
σ -phase F 52 D8b 64.2a/o [V] [60]
Cr-Be CrBe12 F 50 0.2 [61]
Cr-Mn (α-Cr, Mn) AF 300–800 [62]
Cr-Fe σ -phase F 9–47 0.04–0.13 D8b weak itiner. F 49–43a/o [Cr] [58]
(α-Cr, Fe) F 1045–0 2.2–0 Complete solubility if quenched. Initial increase of TC <RT @ 70% [Cr] and [57]
J. P. Liu et al.

continuous decrease of moment


Cr-Co (Cr, α-Co) F Steep decrease of TC with Cr <RT @ 20% [Cr] [57]
σ -phase AF/F TN 300 TC 80 D8b 56.5 a/o [Cr] [63]
Cr-Ni (Ni, Cr) F Steep decrease of TC with Cr <RT 7% [Cr] [47]
Cr-Ir CrIr3 F 425 [64]
Cr-Pt CrPt3 FI 1173 @ 52a/o Pt0 @ 83a/o Pt 2.3/Cr–0.3/Pt [65]
CrPt AF >1573 2.2/Cr [65]
Cr-B CrB2 AF 88 Weak itiner. AF [66]
Cr2 B P F upon slight Fe doping [67]
Cr-Al CrAl7 AF 220 [68]
15 Metallic Magnetic Materials

Cr2 Al AF 670 0.95/Cr [69]


Cr-Ga (Cr,Ga) AF 320–720 ≤10% Ga addition increases TN [70]
Cr-Si CrSi2 Diamagnetic crystals vs F nanowires [71, 72]
Cr-Ge Cr11 Ge19 F 88 [73]
Cr-N CrN AF 286 Magnetostructural transition of first order [74]
Cr-As CrAs AF 265 1.7/Cr Superconducting 2 K @ 8 kBar [75]
Cr3 As2 FI 238 0.58 [76]
Cr2 As AF 393 1.3/Cr [77]
Cr-Sb CrSb AF 723 2.7/Cr Anomalous thermal expansion [78]
Cr-Te Cr2 Te3 F 170 [79]
Cr1−δ Te F 325–360 3/Cr 0<δ<0.25 NiAs type [79]
Mn-Co Co,Mn F 130/10 Rapid drop of TC with Mn concentration; critical composition [80, 81]
35% Mn
MnCo AF 400 [81]
Mn-Ni MnNi3 F 750 Dependent on degree of order [82]
MnNi AF 1070 L10 type [83]
(Continued)
707
708

Table 4 (Continued)
System Compound Type of order TC /TN (K) μ(μB ) Comment Ref.
Mn-Zn MnZn F ∼500 K 0.7 Doping with Ga enhances F moment [84]
Mn-Ru (Mn,Ru) AF <300 Strong concentration dependence; also spin glass [85]
Mn-Rh Mn3 Rh AF 855 [83]
Mn-Pd Mn(Pd,Pt) AF 870 [83]
Mn2 Pd3 AF 640 Superstructure of double CuAu type slightly changes TN [86]
Mn-Ir Mn3 Ir AF 960 [83]
MnIr AF 1145 [83]
Mn-Pt MnPt3 F 400 3.6/Mn 0.4/Pt TC strongly composition dependent [87]
MnPt AF 975 4.3/Mn [87]
Mn3 Pt AF 475 3.3/Mn [87]
Mn-B MnB2 F 157 0.25 [88]
Mn3 B4 AF 392 2.92/2c; 0.44/4g T < 262 K spiral structure [89]
MnB F 578–572 1.92–1.84 [90]
Mn-Al MnAl F 656 1.65 L10 type [91]
Mn-Ga Mn3 Ga2 F 658–748 0.5–0.9 60–67% Mn L10 tetragonal low-temperature phase [92]
(high-coercivity phase)
Mn8 Ga5 F 210 1 62% Mn 973 K annealed and quenched cubic γ brass [93]
Mn2 Ga FI 470 70–72% Mn DO19 hexagonal high temp phase [94]
Mn3 Ga AF/FI 600/800 Magnetostructural phase transitions at 600 and 800 K [95]
Mn-In Mn3 In FI 80 Spin-glass state? [96]
Mn-C Mn23 C6 AF 105 [97]
Mn-Si MnSi H 29.5 0.4 Helimagnet; skyrmions [98]
Mn-Ge Mn11 Ge8 AF/F 140/274 Low T AF spin flop at 140 K to F [99]
Mn5 Ge3 F 300 [100]
Mn5 Ge2 AF/F 315/470110/204 Two superstructures with two TC and first-order spin reorientation [101]
J. P. Liu et al.

transitions; AF ground state


Mn3.25 Ge F 850 K [102]
Mn-Sn MnSn2 AF 324 0.86/Mn [103]
Mn3 Sn AF 420 Spin reorientation 270 K [104]
Mn3 Sn2 F 262 3.0/8d; 2.3/4c Below 227 K canted F of 8d sites 197 K spin reorientation [105]
Mn11 Sn3 AF 423 [103]
Mn-P MnP F 291.5 1.3 Spin reorientation to helix ∼ 50 K [106]
Mn2 P AF 103 0.63/Mn [107]
Mn3 P AF 150 [108]
Mn-As MnAs F 318 3.4/Mn 307 K on cooling; giant MCE [109]
15 Metallic Magnetic Materials

Mn2 As AF 570 Two AF sublattices [110]


Mn-Sb MnSb F 590 3.5/Mn [111]
Mn2 Sb F 550 Spin inversion on doping both on Mn and Sb sublattice [112]
Mn-Se MnSe2 AF 53 Semiconductor [113]
Fe-Co (Fe, Co) F 1043–1388 1.7–2.5 Complete solubility; between 30 and 70% Co; TC coincides with Tα-γ [57]
Fe-Ni FeNi3 F 934 1.2 Superstructure [114]
FeNi F 683 1.7 γ -phase [114]
Fe3 Ni F 800 α-phase [114]
Fe-Cu (α-Cu,Fe) F 1030 2.1 Low solubility of Cu in BCC Fe [57]
Fe-Zr FeZr2 F >300 [49]
Fe2 Zr F 610–800 1.3–1.7/Fe [54]
Fe-Nb Fe2 Nb SDW 10 0.035 Fe-rich TC < 60 K [115]
Quantum criticality
Fe-Mo α-Fe,Mo F 1020 8% [Mo] [57]
Fe-Ru α-Fe,Ru F Slight increase of TC @ 2.2% and then reduction [57]
(Continued)
709
710

Table 4 (Continued)
System Compound Type of order TC /TN (K) μ(μB ) Comment Ref.
Fe-Rh FeRh F-AF RT AF 3.3/FeF3.1/Fe; 1.0/Rh Ordering temperature quite sensitive to slight deviations [116]
from stoichiometry; large MCE
FePd F 760 L10 superstructure; hard magnetic [117]
Fe-Hf Fe2 Hf F 427 1.52/Fe C15 [118]
F 588 C14 [118]
Fe2 (Hf,Ta) F RT C14, exchange inversion [119]
Fe-Os α-Fe,Os F Strong reduction of TC [57]
Fe-Ir α-Fe,Ir F Strong reduction of TC [57]
Fe-Pt FePt3 AF 170 3.3/Fe [120]
FePt F 750 2.8/Fe;−0.25/Pt Hard/soft magnet order/disorder, nanocomposites [121]
Fe3 Pt F 280/430 2.7/Fe;0.5/Pt Disordered/ordered; Invar [122]
Fe-B FeB F 590 1.1 [90]
Fe2 B F 1013 1.91 [123]
Fe-Al α-Fe,Al F Decrease of TC with Al; <RT @ 16%[Al] [57]
Fe3 Al F 823 2.2/FeI1.5/FeII [124, 125]
Fe-Ga α-Fe,Ga F 19a/o Ga magnetostrictive material [126]
Fe3 Ga4 AF/F 360/62 Metamagnet [127]
AF 390 Four inequivalent Fe sites [128]
Fe3 Ga F >870 FCC; F L12 phase transforms to NM DO19 @ high T [129]
F 600 2.5/FeI1.7/FeII BCC; Fe3 Al type [125, 129]
Fe-C Fe3 C F 483 [83]
Fe-Si FeSi F 15 Crossover in magnetic properties, susceptibility max. [130]
above RT
(Fe,Co)Si H Helimagnet and skyrmions [131]
Fe3 Si F 800 K 2.4/FeI1.2/FeII [124]
J. P. Liu et al.
Fe-Ge FeGe2 AF 190 2.4/Fe Two-sublattice ferrimagnet due to vacancies on one sublattice [132]
FeGe H 279 1.0 Cubic B20; skyrmions [133]
AF 340 1.7, 0.5, <0.4 Monoclinic; three different magnetic moments; simple AF above 120 K; [134]
incommensurate at low T
AF 411 1.7 Hexagonal B35 [135]
Fe1.67 Ge F 485 1.6 Ni2 In type; two Fe sublattices S=1 and S= ½, respectively [132]
Fe3 Ge F 740/640 2.2/2.0/Fe Cubic/hexagonal [136]
Fe-Sn FeSn2 AF 378 1.7/Fe Tt = 93 K [137]
FeSn AF 370 1.5/Fe [138]
15 Metallic Magnetic Materials

Fe3 Sn2 F 657 2.0/Fe [139]


Fe5 Sn3 F 620 1.9/Fe [140]
Fe3 Sn F 725 2.3/Fe [140]
Fe-N Fe2 N F 9 [83]
Fe3 N F 567 [83]
Fe4 N F 769 2.2/Fe [83]
Fe-P FeP H 125 0.4 [141]
Fe2 P F 216 1/3f ; 2/3g First-order phase transition; giant MCE [87, 142]
Fe3 P F 716 1.8/Fe [143]
Fe-As FeAs2
FeAs AF 77 K 0.5/Fe [144]
Fe2 As AF 353 1.3/2.1/Fe Two iron sublattices [145]
Fe-Sb FeSb2 P Low-spin high-spin semiconductor [146]
FeSb AF 232 0.9/Fe Excess Fe for stability of phase [147]
Co-Ni Co50 Ni50 F ∼1100–1130 ∼1.2 Sol. soln [57]
Co-Zn CoZn F ∼440 [57]
(Continued)
711
712

Table 4 (Continued)
System Compound Type of order TC /TN (K) μ(μB ) Comment Ref.
Co-Y CoY3 F <2 ∼0 [148]
Co3 Y4 F <2 0.03/Co 43 a/o [Co] [148]
Co2 Y P Metamagnet [148, 149]
Co3 Y F 306 0.55/Co [148]
Co7 Y2 F 639 1.06/Co [148]
F 987 1.50/Co [148]
F 1186 1.62/Co [148]
Co-Zr Co2 Zr F 55 0.2/Co Co2.8 Zr [150]
F 160 0.4/Co Co3.0 Zr [150]
Co4 Zr F 760 [57]
Co23 Zr6 F 766 1.4/Co [151]
Co11 Zr2 F 769 1.24/Co Co5.1 Zr [152]
Co-Nb Co3 Nb ? ? 0.17/Co Probably impurity phase [153]
Co-Ru Co80 Ru20 F ∼800 HCP sol. soln [154]
Co94 Ru6 F 1167 1.47 FCC sol. soln [154]
Co-Rh Co70 Rh30 F ∼990 HCP sol. soln [154]
Co93 Rh7 F 1277 1.61 FCC sol. soln [154]
Co60 Pd4 0 F ∼1080 HCP sol. soln [154]
Co73 Pd27 F 1205 1.49 FCC sol. soln [154]
Co-La Co19 La5 F 616 0.1/Co [148]
Co5 La F 840 [155]
Co13 La F 1297 1.58/Co [148]
Co-Hf Co3 Hf F 40 0.12/Co [153]
Co23 Hf6 F 499 0.66/Co [153]
J. P. Liu et al.
Co7 Hf F 600 1.14/Co [153]
Co-Re Co50 Re50 F Sol. soln; no magnetic data found [156]
Co-Os Co80 Os20 F ∼630 HCP sol. soln [154]
Co99 Os1 F 1258 1.66 FCC sol. soln [154]
Co-Ir Co70 Ir30 F ∼680 HCP sol. soln [154]
Co90 Ir10 F 1108 1.43 FCC sol. soln [154]
Co-Pt CoPt3 F 460 1.29/Co; 0.45/Pt Stoichiom.; disordered [157]
∼430 22 a/o [Co]; disordered [158]
F 320 1.64/Co; 0.26/Pt Stoichiom.; ordered [157]
15 Metallic Magnetic Materials

∼330 22 a/o [Co]; ordered [158]


CoPt F 750 1.6/Co;0.25/Pt [149]
F ∼810 Stoichiom.; disordered [158, 159]
F ∼710 Stoichiom.; ordered [158, 159]
Co3 Pt F 1190 L12 [83]
Co85 Pt15 F 1261 1.57 FCC sol. soln [154]
Co-Ga CoGa P F for >52 a/o [Co] [149]
Co-As Co2 As F ∼12 hex [160]
30 0.05/Co Taken from Bec91 [161]
Co-S CoS2 F 124 0.84/Co [113]
Co-Se CoSe2 AF 93 [113]
Ni-Cu Ni,Cu F Sol soln. 30% [Cu] TC <RT [114]
Ni-Y Ni3 Y F 33 0.05/ Taken from Bus77 [162]
F 30 0.004 Weak itinerant F [163]
Ni7 Y2 F 58 0.03 Taken from Bus77 [164]
Ni17 Y2 F 160 0.32/Ni [148]
(Continued)
713
Table 4 (Continued)
714

System Compound Type of order TC /TN (K) μ(μB ) Comment Ref.


F 149 0.19/Ni [165]
Ni-Zr Ni5 Zr P Exchange enhanced [156]
Ni-Ru Ni88 Ru12 F 266 0.23 Taken from Cra60; [166]
FCC sol. soln
Ni-Rh Ni70 Rh30 F 141 0.24 FCC sol. soln [154]
Ni90 Rh10 F 511 0.61 FCC sol. soln [154]
Ni-Pd Ni31 Pd69 F 0.48 Taken from Cra60; [166]
FCC sol. soln
Ni73 Pd27 F 0.59 Taken from Cra60; [166]
FCC sol. soln
Ni-La Ni3 La P [148]
P Metamagnetic ≤50 K [167]
Ni7 La2 AF 54 Hexagonal [168]
F ∼65 Rhombohedral [167]
Ni-Os Ni95 Os65 F 415 0.38 FCC sol. soln [154]
Ni-Ir Ni94 Ir6 F 461 0.51 FCC sol. soln [154]
Ni-Pt NiPt F 95 0.23 Stoichiom.; disordered [169]
F 0.28/Ni, 0.17/Pt Stoichiom.; disordered [170]
F ∼100 Stoichiom.; disordered [158]
P Stoichiom.; ordered [158, 169]
Ni55 Pt45 F 202 0.25 Taken from Cra60; [171]
FCC sol. soln
Ni3 Pt F ∼370 Stoichiom.; disordered [158]
F ∼320 Stoichiom.; ordered [158]
Ni-Au Ni70 Au30 F 250 0.27 [57]
Ni-Al Ni3 Al F 41 [172]
Ni-Ga Ni3 Ga P F for ≥76 a/o [Ni] [172]
J. P. Liu et al.
Table 5 Martensitic transformation temperature, T0 ; Curie temperature, TC ; the martensitic structure type and the field-induced strain for Ni-Mn-Ga single
crystals. NM, non-modulated martensite; 5M, five-layer stacked martensite; 7M, seven-layer stacked martensite
Martensitic
Composition T0 (K) TC (K) structure type Field-induced strain Remarks Ref.
Ni50 Mn25 Ga25 276 – NM 0.2% [176]
Ni52 Mn22.2 Ga25.8 298 – – 0.31% [177]
Ni52 Mn23 Ga25 311 – – 0.27% [178]
15 Metallic Magnetic Materials

Ni49.8 Mn28.5 Ga21.7 318 368 5M 6% 1 MPa bias stress [179]


Ni48 Mn30 Ga22 308 – NM 5% Single-variant sample [180]
Ni48 Mn31 Ga21 300 371 NM 5.1% 0.27 MPa bias stress [181]
Ni52 Mn24.5 Ga23.5 287 – – 5% Intermartensitic transformation at [182]
240 K
1.2% In the modulated martensite
Ni48.8 Mn29.7 Ga21.5 340 368 7M 9.5% [183]
Ni51 Mn24 Ga25 235 – – 1.1% Temperature varies under [184]
1.2 T bias field
Ni50 Mn27.5 Ga22.5 298 370 5M 6.3% Pre-compressed single variant [185]
Ni48.4 Mn27 Ga24.6 178 382 5M 3% Reversible; 2.7 MPa bias stress [186]
Ni50 Mn28.5 Ga21.5 320 376 5M 6.2% – [187]
Ni50 Mn29.2 Ga20.8 339 368 5M 7.1% At room temperature [188]
6% Around martensitic transformation
Ni50.8 Mn28.4 Ga20.8 348 369 7M + 5M 7.3% At room temperature [189]
4.5% Around martensitic transformation
715
716

Table 6 Martensitic transformation temperature, T0 ; Curie temperature, TC ; the martensitic structure type and the field-induced strain for Ni-Mn-Ga
polycrystals, composites, and thin films
Composition T0 (K) TC (K) Martensitic structure type Field-induced strain Remarks Ref.
Ni50.9 Mn27.1 Ga22 328 371 5M 0.1% Magneto-thermal training [190, 191]
epoxy resin
Ni50.9 Mn27.1 Ga22 1% – [192]
polycrystalline fibers
Ni50.6 Mn28 Ga21.4 – – – 0.115% Introducing pores with similar sizes [193]
polycrystalline foam to grain size
8.7% Introducing pores smaller than the [194]
grain size
Ni51 Mn29 Ga21 328 – 5M 0.16% Free strain [195]
polycrystalline
1% Polyurethane fills the cracks; after [196]
thermomagnetic training
Ni50 Mn29 Ga21 polycrystalline 328 – 5M 1% – [197]
Ni50.5 Mn24 Ga25.5 226 – – 1.5% Temperature varies under [198]
polycrystalline 1.5 T bias field
Ni51.4 Mn28.3 Ga20.3 311 – 5M 0.8% [199]
film
J. P. Liu et al.
15 Metallic Magnetic Materials 717

can be realized . Data on these non-Ni-Mn-Ga single crystals and polycrystals are
listed in Tables 7 and 8, respectively.

Magnetic Heusler Compounds

Magnetic Heusler compounds exhibit unique and tunable magnetic properties


because of their particular electronic structure [229]. This materials class, named
after the German chemist Friedrich Heusler, is distinguishable from regular mag-
netic alloys, in the sense that they are ferromagnetic intermetallic compounds with
face-centered cubic crystal structure. In a wide context, the interpretation of the so-
called “half”-Heusler compounds in terms of the chemical Zintl concept is possible
as they may be described as ternary relatives of the classical binary semiconductors
such as GaAs and ZnS. An interpretation as borderline materials between metallic
alloys and covalent semiconductors is justified because they exhibit characteristics
of both materials classes. LiMgAs and TiNiSn are examples of such half-Heusler
semiconductors [230] that crystallize in the C1b structure (prototype MgAgAs,
cF12, F-43m, 216, Fig. 7) [231]. In these compounds with general formula XYZ,
the atoms occupy the three Wyckoff positions 4a (Z), 4b (X), and 4c (Y) of the fcc
lattice, whereas the 4d position stays vacant. In the chemical notation, the X atom
is the most electropositive among the constituting atoms and jointly forms a rock
salt (NaCl)-like sublattice with the Z atoms that reflects the rather ionic bonding
character between X and Z. In contrast, Y atoms and Z atoms constitute a zinc
blende-like lattice, which is more covalently bonded [232]. The band gap of the
lithium compound LiMgAs lies between the bonding and antibonding sp-states (four
occupied bands, eight valence electrons per primitive unit cell and formula unit),
while in case of TiNiSn the band gap is found between the bonding and antibonding
spd-states (18 valence electrons).
Magnetic half-Heusler compounds with C1b structure are found within rare-
earth elements (RE) or manganese-containing materials. In these compounds, the
magnetic moments are commonly found on the X sites. Contributions from the Y
sites are usually small or negligible. Prominent systems are the REPtBi-Heusler
compounds, which were also identified as topological insulators [233, 234]. These
compounds are 18 valence electron compounds under the assumption that the 4f -
electrons are localized at the rare-earth atom and hence do not contribute to the
valence band. The semiconducting rare-earth-containing C1b-Heusler compounds
are antiferromagnetic with low Néel temperatures. Most half-Heusler compounds,
containing rare-earth atoms, may be synthesized with manganese substituting the
rare-earth element, maintaining the magnetic properties. A typical compound is
MnNiSb (or NiMnSb in the notation of physics [235]) for which Mn exhibits
approximately a d4 -configuration with a localized moment of 4 μB . The magnetic
moment of nickel is close to zero. Magnetism in the half-Heusler alloys thus is
induced by the strong magnetic moments of Mn or rare-earth elements on the X
sites and is of a localized character. These atoms on X sites experience a tetrahedral
crystal field, spanned by the Y atoms. The mechanism of the localization of the
718

Table 7 Martensitic transformation temperature, T0 ; Curie temperature, TC ; the martensitic structure type and the field-induced strain for non-Ni-Mn-Ga
single-crystal compounds. 14M, 14-layer stacked martensite
Martensitic Field-induced
Composition T0 (K) TC (K) structure type strain (%) Remarks Ref.
Mn50 Ni25 Ga25 588 270 NM 4.0 Thermal induced Ref[200]
2.0 Magnetostrain
Co45 Ni25 Ga20 351 – NM 1.0 Oriented polycrystalline [201]
Ni51 Fe18 Ga27 Co3 338 300 14M 0.7 At room temperature [202]
Ni45 Co5 Mn36.6 In13.4 292 382 14M 2.9 3% pre-strain at room temperature [203]
Ni45 Co5 Mn36.5 In13.5 260 400 – 3.13 Compressive stress 125 MPa, [100] single crystal [204]
Ni49 Fe18 Ga27 Co6 – – NM 8.5 Pre-strain (8 MPa), single variant at room temperature [205]
Ni49.5 Fe14.5 Mn4 Ga26 Co6 320 420 – 11.3 Pre-strain (8 MPa), single crystal at room temperature, [206]
variant rearrangement
Fe43 Mn28 Ga29 213 – NM 0.6 Volume change +1.35% [207]
Fe70 Pd30 293 FCT 0.6 Single crystal, 256 K, axial field [208, 209]
2.3 kOe/978-3-030-63208-3/[100], applying field in
(100) plane, reversible, variant rearrangement
J. P. Liu et al.
Fe68.8 Pd31.2 293 FCT 3.1 Single variant, 77 K, [100], variant rearrangement [210]
Fe3 Pt 85 FCT 2.3 Single variant, 4.2 K, [100], variant rearrangement [210][211]
Fe3 Pt 85 FCT Recoverable, 0.6 Single crystal, [001], 4.2 K, capacitance method in a pulsed [212]
magnetic field
1.7
Fe3 Pt 85 FCT Recoverable, 0.45 Single crystal, [001], 20 K [213]
Recoverable, 0.95
Mn0.965 CoGe 292 – Orthorhombic 11.6 (c-axis Derived from XRD analysis; volume change +4% [213]
expansion)
15 Metallic Magnetic Materials

Mn0.84 Fe0.16 NiGe 241 – Orthorhombic 12.3 (c-axis Derived from XRD analysis; volume change +2.7% [214]
expansion)
Mn0.9 Ni0.1 CoGe 290 – Orthorhombic 11.3 (c-axis Derived from XRD analysis; volume change +3.9% [215]
expansion)
Mn0.64 Fe0.36 NiGe0.5 Si0.5 348 – Orthorhombic 12.6 (c-axis Derived from XRD analysis; volume change +3.6% [216]
expansion)
719
720

Table 8 Martensitic transformation temperature, T0 ; Curie temperature, TC ; the martensitic structure type and the field-induced strain for non-Ni-Mn-Ga
polycrystalline compounds. 10M, ten-layer stacked martensite; 6M, six-layer stacked martensite
Field-induced strain
Composition T0 (K) TC (K) Martensitic structure type (%) Remarks Ref.
Ni53 Mn25 Al22 234 310 NM 0.1 [217]
Ni50 Fe25 Ga25 142 430 NM 0.3 Ribbon [218]
Ni50 Mn34 In16 305 210 10M 0.12 [219]
Co50 Ni20 Ga20 270 >400 NM 0.15 Ribbon [220]
Ni45 Co5 Mn37 In13 320 390 NM 0.25 [221]
Ni50 Mn33 Fe4 In13 140 320 10M 0.28 [222]
Ni43 Co7 Mn39 Sn11 305 400 6M + 10M 1.0 1.3% pre-strain at 310 K [223]
Ni49 Fe1 Mn36 Sn14 170 322 NM 0.2 Thermal induced [224]
Ni50 Mn35 Sn15 230 – – 0.2 Thermal induced [225]
Ni46 Cu4 Mn38 Sn12 280 – – 0.12 [226]
Fe50 Mn22.5 Ga27.5 132 – NM 3.6 Thermal induced, volume [227]
change +0.85%
Ni35 Co15 Mn35 Ti15 255 391 5M 0.25 Volume change −1.95% [228]
J. P. Liu et al.
15 Metallic Magnetic Materials 721

Fig. 7 (a) Zinc-blende crystal structure: red spheres, Ga; green spheres, As. (b) Half-Heusler
structure: red spheres, Mg (Ni); blue spheres, Li (Zr); green spheres, As (Sn). (c) Schematic density
of states of ZrNiSn: green, bonding and antibonding sp3 -states; red, non-bonding d10 -states. (d)
Schematic density of states of half-metallic NiMnSb (see Refs. [236, 237])

Mn moments differs from the localization 4f -electrons in rare-earth atoms, which


are completely excluded from the valence bands, in contrast to the Mn 3d-states. It
is interesting to note that NiMnSb and its relatives are half-metallic ferromagnets
[235]. Half-metallic ferromagnets are metallic for the majority states and exhibit a
band gap in the minority density of states at the Fermi level.
Heusler alloys are compounds with the chemical formula X2 YZ or X Y Z
wherein X, X , and X are transition metals, Y is a transition metal or a lanthanide
(rare-earth metal), and Z is a main group element. Regular Heusler compounds
crystallize in the L21 structure (prototype: Cu2 MnAl, cF16, Fm-3m, 225) [231]
where the elements are placed on the 8c (X), 4b (Y), and 4a (Z) Wyckoff positions
of the fcc lattice. In L21 Heusler materials, the crystal field at the Y position evolves
722 J. P. Liu et al.

to an octahedral field due to the filling of the voids at the 4d position of the C1b
structure.
Many compounds (≈1000) are known in this family [230]. Due to the complex
interplay of atomic interactions, a variety of magnetic phenomena have been
observed, of which ferromagnetism, ferrimagnetism, or antiferromagnetism are the
most intensively investigated. In case of X2 YZ Heusler compounds, semiconductors
are found with 8, 18, or 24 valence electrons per formula unit such as Fe2 Val [238].
If the number of valence electrons differs from 24, these materials are usually ferro-
or ferrimagnetic, where the magnetic moment per formula unit is directly related
to the number of valence electrons minus 24 resulting in a Slater–Pauling type
behavior [239, 240]. Compounds containing Co, Mn, or Fe atoms on the X and/or
Y sites usually follow the Slater–Pauling rule. Prominent examples are Co2 MnSi
and Co2 FeSi that exhibit a strong localized moment at Mn or Fe and an itinerant
moment at Co. The local moments sum to 5 or 6 μB , respectively. Another out of
many possible examples is ferromagnetic Co2 TiSn where the magnetic moments
mainly reside on the Co sites [241]. The compound has a magnetic moment
of 2 μB , and a Curie temperature of ≈ 350 K is observed. Co2 MnSi performs
best of all Heusler compounds in magnetic tunneling devices. Fe-doped Co2 MnSi
electrodes exhibited giant tunnel magnetoresistance (TMR) ratios of 2610% at
4.2 K and 429% at 290 K [242]. Co2 MnSi was subject of a detailed investigation
accessing its surface spin polarization which has been shown to be certainly half-
metallic within the accuracy of spin-polarized photoelectron spectroscopy [243].
For ferromagnetic, half-metallic Heusler compounds, a general relation between
valence electron count NV and Curie temperature was established according to
which TC scales approximately linearly by 175 K per electron deviating from
NV = 24 [244]. Co2 FeSi is ferromagnetic with the Fermi energy being located at
the upper edge of the minority band gap. Therefore, it is at the borderline between
a regular and a half-metallic ferromagnet. It exhibits the highest Curie temperature
of all known Heusler compounds with TC ≈ 1120 K [245].
In Heusler compounds, the X sites are located in a hetero-cube (tetrahedral
symmetry) that is formed by the Y and the Z atoms (Fig. 8). Within a unit
cell, eight such sites exist. The electronic states of the X atoms develop to a
delocalized magnetic sublattice in comparison to the localized moment of Mn at
the Y site. The presence of two strong magnetic sublattices on the X and Y sites may
lead to complex magnetic structures such as ferrimagnetic order. For example, in
hypothetically cubic Mn3 Ga, the moments on the X and Y sites are expected to be
coupled in an antiparallel manner [246]. Similar to the ferromagnetic systems, in
ferrimagnetic half-metallic Heusler compounds, the Curie temperature scales with
the sum of the local magnetic moments on the two antiparallelly aligned sublattices
[247].
The Wyckoff position 4b is most often occupied by the most electropositive
element (i.e., earlier transition metal). The 8c position of space group 225 can be
split into the positions 4c and 4d of space group 216 keeping the fcc lattice. In many
Mn2 -based Heusler compounds, an arrangement is observed where half of the Mn
atoms occupies site 4b and the other half is found on 4d of space group 216. This
15 Metallic Magnetic Materials 723

a b 6
4

E-εF [eV]
2
0
-2
-4

Energy
-6
-8
-10
-12

c d
Magnetic moment per atom m [µ8]

2.5 bcc fcc


localised Fe itinerant Meas. Tc 1200

C
Co2FeGa

o
o

Fe
Tc

2
FeCo Calc.

M
Co2TM

Si
nS
2.0 FeNi 1000

i
FeCo

Curie Temperature Tc [K]


NiCo
FeCr Co

C
NiCu

o
FeV 800

M
2
1.5 Co2FeSi

nS
CoCr

n
o
NiCr

C
2
NiMnSb Co2MnSi

rG

C
600

o
a

M
2
1.0
C

nA
CoMnSb Co2MnAl
o
VG
2

l
a

Co2CrGa 400

C
Ni

o
0.5

C
2
Co2VAI

rA
l
GGA 200
Co2TiAl
0.0 Cr Co2TiAl Co2VSn
0
6 7 8 9 10 11 25 26 27 28 29 30
Valence electrons per atom Nv. Valence electrons per f.u. Nv.

Fig. 8 (a) L21 Heusler structure, (b) band structure of Co2 MnSi, (c) Slater–Pauling rule, and (d)
scaling of the Curie temperature in Co2 YZ Heusler compounds (see references [236, 237])

situation leads to the inverse Heusler structure with the lower symmetry of space
group 216 (Fig. 9). This space group does not contain an inversion center, and it has
been argued that the Dzyaloshinskii–Moriya interaction may play a crucial role on
the magnetic properties. Most prominent examples of the ferrimagnetic materials in
the inverse structure are Mn-rich Heusler compounds [247–249]. As already stated
by Kübler in 1983, Mn plays the unique role in Heusler compounds, due to the
localization of its site moment, that is traced back to a strong exchange split [250].
An example for an inverse Heusler compound is (MnCo)MnAl [251]. Mn occupying
the 4b site shows a localized moment of ≈ 3 μB . The moment of the atoms in the
Mn-Co plane (Mn(4d) and Co(4c)) counterbalances the strong local moment by
moments of mMn(4d) = −2 and mCo(4c) =1 μB , resulting in a total moment of 2 μB
according to the Slater–Pauling rule: Mtot = NV − 24 = 7(Mn4d) + 9(Co4c) +
7(Mn4b) + 3(Ga4a) − 24.
Contrary to expectations, D03 -type ordered cubic Mn3 Ga could not be obtained
as stable bulk material from chemical synthesis. Hexagonal antiferromagnetic
Mn3 Ga also turned out to be metastable. Surprisingly though, a stable and tech-
nologically useful tetragonal, ferrimagnetic structure exists (prototype: TiAl3 , tI8,
I4/mmm, 139). Tetragonal magnetic Heusler compounds are of special interest,
thanks to their potential for spin transfer torque magnetic random access (STT-
MRAM) devices [24, 252], and they have been considered as candidate materials
724 J. P. Liu et al.

Fig. 9 Crystal structure of (a) the magnetic Heusler compound Mn2 VAl, (b) the inverse Heusler
compound Mn2 CoGa

replacing the nowadays widely used rare-earth-based permanent magnets [253].


Several attempts on the explanation of the occurrence of the tetragonal distortion are
given in the literature. They cover different perspectives on the issue. The tetragonal
instability in momentum space is interpreted in terms of Fermi surface nesting,
while a real space interpretation is given by soft modes in the phonon spectrum
appearing for certain q-vectors of acoustic vibrations. Another interpretation is
based on the so-called band Jahn–Teller effect , the crystal analogue to the atomic
Jahn–Teller effect in the crystal field theory of complexes/molecular species. In
molecular complexes, partially occupied degenerate electronic states may cause
a distortion of the surrounding coordination sphere, as it is often found for Mn-
ions with d4 configuration in an octahedral coordination. Similarly, the tetragonal
instability in Mn-containing Heusler compounds is often approached with respect to
the band Jahn–Teller effect, due a Mn atom occupying the Wyckoff position 4b, that
is octahedrally coordinated by the 8c position, for example, in Ni2 MnGa or Mn3 Ga.
The tetragonal distortion on the Y site results in distorted local environment of the
Z atom as well. Thus, if there is Si on the Z site, the tetragonal deformation of the
unit cell is strongly suppressed due to Si preferring a highly symmetric tetrahedral
environment as to optimize its bonding situation as sp-hybridized species. Typically,
compounds with Al, Ga, Ge, Sn, Pb, and Sb on the Z site allow for a tetragonal
distortion. On the Y site, the moment is typically localized: the d4 configuration is a
Jahn–Teller ion only in the octahedral environment. The octahedron may expand or
shrink along one axis. One example of a Jahn–Teller-driven tetragonally distorted
Heusler compound is Mn3 Ga [254]. It has been shown on the examples of Mn3 Ga,
Mn2 FeGa, and Mn2 NiGa that the peak of the DOS at the Fermi edge mainly arises
from atoms of the 8c position (or 4c/4d in inverse Heusler compounds). A second
reason why Heusler or half-Heusler compounds may exhibit hexagonal or tetragonal
distortions is the occurrence of particular van Hove singularities in the electronic
15 Metallic Magnetic Materials 725

band structure. In general, such singularities may additionally lead to a very high,
peaked density of states at the Fermi energy. One particular van Hove singularity is
associated with a saddle point in the dispersion curves in the vicinity of the Fermi
energy. It appears sometimes very pronounced in the direction of the fcc Brillouin
zone. It often causes less pronounced maxima in the density of states compared to
other global maxima. The occurrence of the saddle point-type van Hove singularity
in Heusler compounds is often associated with the band Jahn–Teller effect. Similar
to the removal of a high density of states at the Fermi energy, the van Hove
singularity is removed or shifted away from the Fermi after a structural distortion.
The distortion will open a gap or more often a pseudo-gap in the band structure
to lower the total energy of the system by lowering the band energy. A pseudo-
gap describes the state in which the band structure may not be completely gapped
after the distortion. Examples of known tetragonally distorted Heusler compounds
are the magnetic shape memory compounds: Ni2 MnGa or Mn2 NiGa [200, 255]. At
increased temperature or in a distinct magnetic field, these materials are cubic. It is
found that the origin of the tetragonal distortion at low temperature is a localization
of a van Hove singularity near to the Fermi energy. Following this observation,
a series of new compounds such as Mn3−x Fex Ga (0 ≤ x ≤ 2), Mn2−x Co1+x Ga
(0 ≤ x ≤ 0.5) was synthesized where such a van Hove singularity is responsible
for the tetragonal distortion. In rhodium-containing Heusler compounds such as
Rh2 FeSn, a tetragonal distortion also occurs because of a van Hove singularity
in the vicinity of the Fermi energy. This tetragonal distortion was attributed to an
electronic instability of the band Jahn–Teller type by Suits [256].
The tetragonal distortion is further influenced by a strong spin–orbit interaction.
This is the case for 4d or 5d transition metals on the tetrahedral (8c) sites. Known
compounds are, for example, Rh2 FeSn, Rh2 CoSn, and Rh2 FeSb [256]. The splitting
of the d states into d3/2 and d5/2 will have an additional impact on the states at
the Fermi energy that are responsible for the band Jahn–Teller effect. Furthermore,
the spin–orbit interaction results in a pronounced magnetocrystalline anisotropy in
tetragonal Heusler compounds.
The formation of new alloys using two known Heusler compounds offers a
possibility to create new alloys and tune their properties such as the structural
parameters, as well as magnetic characteristics or electric transport quantities. An
example for mixing of similar compounds is the series of alloys between the
cubic systems Mn2 CoGa and tetragonal Mn3 Ga [246, 251, 254]. Tetragonal, hard
magnetic Mn3 Ga was doped in thin films with Co which caused magnetic moment
to drop and the coercive field to increase [251]. The systems with a Mn content
of x ≤ 0.5 remain in the tetragonal lattice but become cubic for x > 0.5 [249].
The magnetic moments of both cubic and tetragonal compounds vary linearly with
respect to the valence electron count. This is the famous Slater–Pauling rule for
the cubic compounds (x > 0.5) and a pseudo-Slater–Pauling-type behavior for the
tetragonal alloys as shown in Fig. 10 and Ref. [249].
In case of cubic compounds, this means that the magnetic moment per formula
unit is exactly the number of valence electrons minus 24. The Slater–Pauling-type
behavior is often a sign of bulk half-metallicity (100% spin polarization) in that
726 J. P. Liu et al.

Fig. 10 (a) Structure of a tetragonal inverse Heusler compound and corresponding behavior of
the calculated spontaneous magnetization due to the tetragonal distortion of the (b) Mn2 Y3d Ga,
(c) Mn2 Y4d Ga, and (d) Mn2 Y5d Ga compounds

varying the number of valence electrons varies the filling of the metallic spin-
polarized band such that the moment is increased or decreased directly proportional
to the number of valence electrons. This is true as long as the Fermi energy
remains within the energy gap of the semiconducting spin-polarized band. These
experimental observations are in agreement with the finding that a tetragonal
distortion occurs when a high density of states appears in the vicinity of the
Fermi energy. Cubic Mn3 Ga and tetragonal Mn3 Ga doped with Co are only two
possibilities. The knowledge about the Slater–Pauling rule helps to identify other
compositions such as the 1:1 combination of Fe2 MnAl (26 valence electrons) and
Mn2 VAl (22 valence electrons). Deviations of the 24-electron rule are possible in
case of tetragonal, inverse Heusler compounds [249]. Work on Ru-doped Mn2 Ga
shows that crossing the compensation point, where the anomalous Hall effect
chances sign, is possible [257].
Fully compensated half-metallic ferrimagnets are very interesting materials
[246]. An overview, predicting a large number of half-metallic fully compensated
ferrimagnetic Heusler compounds, is given in Ref. [258]. In the system Mn-Pt-
Ga, a new tetragonal Heusler compound with a compensated ferrimagnetic state
was designed using density functional theory [259], explicitly treating Mn-Pt
substitution. In the vicinity of the compensation point xcomp , a giant exchange
bias of more than 3 T and a large coercivity were established. The large exchange
15 Metallic Magnetic Materials 727

anisotropy originates from the exchange interaction between the compensated host
and ferrimagnetic clusters that arise from intrinsic anti-site disorder [259].
In magnetic Heusler compounds, the alignment of adjacent spins often is
assumed to be collinear, which has been proved correct in many systems, especially
for the soft magnetic Co2 YZ alloys. With the observation of spin canting and
reorientation in Mn2 RhSn and the persistent interest in topological properties,
noncollinear order found its way back into the focus of Heusler compound research.
Spin reorientation from collinear to canted spins was observed for Fe2 MnSi [260,
261], and a conical ground state has even been suggested for thin films of Mn3 Ga.
The study of some other compounds is enlightening [260, 261]. A phase transition
in the magnetization in Mn2 RhSn could be resolved with the help of density
functional theory [262]. The phase transition could be modeled through a spin
reorientation transition from a noncollinear to a collinear ferrimagnetic structure.
Mn2 RhSn is only one compound with a small energy difference between the out-
off-plane and in-plane magnetization [249]. The relatively small magnetocrystalline
anisotropy may be counteracted by the Dzyaloshinskii–Moriya exchange in this
non-centrosymmetric structure [262]. Mn2 RhSn exhibits a topological Hall effect
and – similar to related Heusler compounds – shows evidence for skyrmions.
However, we are sure that Mn2 RhSn is not the only Heusler compound revealing
noncollinear magnetic order, and new physics will emerge in this subgroup of
Heusler compounds.

Intermetallic Compounds of 3d-4f Type

The class of intermetallic compounds formed between the rare-earth elements (4f
elements) and the 3d transition metals (Fe, Co, and Ni) contains a huge number of
binary and ternary systems. Here we survey selected systems which are of particular
interests for hard magnetic and magnetocaloric applications.
The R2 M7 (R = rare earth, M = Fe, Co, Ni) may exist in either hexagonal
Ce2 Ni7 -type (P63 /mmc) or rhombohedral Gd2 Co7 -type (R3m) structures. Both
allotropic forms often occur together [263]. Th2 Fe7 is the only reported R2 Fe7 phase
to date. It has a ferromagnetic structure below 570 K and a saturated magnetization
of 9.59 μB /formula unit [264]. For M = Co or Ni, the R2 M7 compounds form for
all the rare-earth elements. The magnetic properties of the R2 M7 compounds are
summarized in Tables 9 and 10 for M = Co, Ni, respectively.
A newly studied stable system R5 M17 was identified in the Nd-Fe binary
diagram [284] with ferromagnetic order and a TC of 503 K, later identified as
Nd5 Fe17 [285] crystallizing in the P63 /mcm space group. Its magnetization is 46.5
and 37.5 μB /formula unit at 4 and 300 K, respectively [286] (see Table 11). This
corresponds to a large magnetic moment per Fe atom of 2.1 μB /f.u but with a
basal plane spin orientation. Nd5 Fe17 and Sm5 Fe17 are the only reported R5 Fe17
phases to date. The Sm isotype is more difficult to synthesize and was first reported
from the investigation of the Sm-Fe-Ti system where partial Ti for Fe substitution
is possible [287] and more recently in a similar manner in the Sm-Fe-V ternary
728

Table 9 Lattice parameters and magnetic properties of R2 Co7 compounds


Easy Anisotropy
magnetization parametersBa (T),
a(10−3 nm) crho (10−3 nm) chex (10−3 nm) TC (K) Ms (μB /f.u.) direction TSR (K) K (MJ/m3 ) Ref.
Y2 Co7 500.2 3669 639 9.6 Ba = 8T [265, 266]
La2 Co7 510.1 3652 2463 479 8.4 at 4 K c K1 = 2 MJ/m3 [267]
6.8 at 300 K 4 K (Ba = 8T at
20 K 6.7T at
300 K)
K1 =1.4 MJ/m3 at
RT
Ce2 Co7 494.0 3652 2447 50 0.9 at 4 K c [268]
Pr2 Co7 506 3652 2443 574 10.5 c [269, 270]
70S6
Nd2 Co7 505.9 3642 2439 613 14.6 a [269–271]
Sm2 Co7 504.1 3631 2433 713 10.1 Cone 15◦ vs c 68 K1 = −5.1 MJ/m3 [269, 270, 272]
then c T > TSR K2 = −37 MJ/m3
4K
Gd2 Co7 502.2 3624 2419 775 3.45 c [269, 270, 273]
Tb2 Co7 500.8 3618 710 7.3 [210] to c above 440 [269, 274]
TSR
Dy2 Co7 499.2 3613 640 [269]
Ho2 Co7 497.9 3612 670 [275]
Er2 Co7 497.3 3611 670 [275]
Tm2 Co7 496.5 3605 640 2.8 Cone to c 45 Tcomp. 80 K [275, 276]
Lu2 Co7 494.6 3598.1 [275]
Th2 Co7 503.0 3691.0 24.62 Para. – [263, 264]
J. P. Liu et al.
Table 10 Lattice parameters and magnetic properties of R2 Ni7 compounds
a(10−3 crho (10−3 chex (10−3 Ms Magnetic
nm) nm) nm) TC (K) (μB /f.u.) anisotropy
SMax (K) θ p (K) Meff (μB /f.u.) Ref.
Y2 Ni7 494.0 3612.0 – 58 0.4 32 8.28 [277, 278]
La2 Ni7 505.8 2471 AF TN = Bmeta = 6.1 T at 70 6.3 [277, 279]
51(hex) 4.2 K (hex)
15 Metallic Magnetic Materials

K > 0.16 MJ/m3


La2 Ni7 505.8 3698.0 F 65 (rh) 50 7.28 [280]
Ce2 Ni7 4.927 2445 48 0.26 48 [278]
Pr2 Ni7 501.5 3636.0 2423 85 4.36 85 [278]
Nd2 Ni7 500.0 3631 2423 87 4.14 87 [278]
Sm2 Ni7 496.9 3653 [278]
Gd2 Ni7 496.0 3614.0 2409 119 12.65 118 [278]
Tb2 Ni7 494.0 3610 2406 101 10.72 101 [278]
Dy2 Ni7 493.5 3608 2405 81 13.30 TSR = 35 K −7 J 81 14.85 [278, 281]
kg−1 K−1
Ho2 Ni7 492.0 3604 2492 70 12.57 TSR = 25 K −12 J 65 16.1 [278, 281]
kg−1 K−1
Er2 Ni7 491.0 3571 2491 67 12.28 67 [278]
Yb2 Ni7 496.5 3586 – [282]
(HP)
Th2 Ni7 494.5 – 2480 [283]
729
730

Table 11 Magnetic properties and room temperature lattice parameters of R5 Fe17 compounds
Easy Ms at
magnetization 300 K
a(10−3 nm) c(10−3 nm) Space group direction TC (K) Ms at 4 K (μB /f.u.) Hc (MA m−1 ) (μB /f.u.) Ref.
Nd5 Fe17 2021.4 1232.9 P63 /mcm Basal plane 503 46.5 – 37.5 [285, 286]
Sm5 Fe17 2052.0 1229.0 P63 /mcm c 550 – 1.9 25 [287, 288, 293]
J. P. Liu et al.
15 Metallic Magnetic Materials 731

system [288]. Alternatively, it can be stabilized by rapid quenching and subsequent


controlled annealing treatment. The Sm for Nd substitution leads to a uniaxial
magnetic orientation along the c-axis. Coercivity as large as i Hc = 5.0 MAm−1 has
been reported at 4 K for the pure Sm5 Fe17 compound as metastable phase obtained
by melt spinning [289, 290] (1.9 MAm−1 at room temperature). Neutron diffraction
measurements have been reported for both Nd5 Fe17 and the partially Sm- or Ti-
substituted compounds [291, 292]. No isotype R5 M17 phase has been reported for
M = Co or Ni.
The RM5 compounds crystallize in the CaCu5 structure type and the P6/mmm
space group. RM5 compounds are formed for the entire rare-earth series with
M = Ni only. In contrast, they are mainly stable for the light rare-earth elements
with M = Co and is only reported for a few R elements with M = Fe. However,
partial substitution of Co pairs for R atoms stabilizes heavy rare-earth R1−δ Co5+δ
compounds up to R = Tm [294] corresponding to final composition of typically
ErCo6 . ThFe5 which orders ferromagnetically at 680 K is a stable compound and
a saturation magnetization of 9.4 μB /formula unit [264, 273], whereas NdFe5
has been only obtained as metastable phase by rapid solidification [294] and is
ferromagnetic up to 370 K. The structural and magnetic properties of the RFe5
compounds are listed in Table 12. The large Co–Co exchange interactions lead to
high ordering temperature for the RCo5 series of compounds as listed in Table 13.
At 293 K, a record uniaxial magnetocrystalline anisotropy of 16 T is obtained for
YCo5 . A similar value has been reported for LaCo5 [295]. This results from a large
anisotropy of K1 = 7.4 MJ/m3 at 4.2 K as has been reported for the Co sublattice
in YCo5 [296, 297]. This added to the Sm anisotropy leads to a very large value
K1 = 26 MJ/m3 for SmCo5 which is consequently an excellent hard magnetic phase
[298]. The magnetic moment carried by the Co atoms in the RCo5 series is about
the same as in elemental cobalt.
Several neutron diffraction experiments have been reported on RCo5 either
powders or single crystals. In addition, it is possible to quantify the orbital and
spin contribution to the magnetic moment on each Co site and determine the wave
function occupancy of the different orbitals. This has been done for YCo5 and other
isotype compounds [301, 302]: the spin contribution to the magnetic form factor has
been determined to be 74% and 84% for the two inequivalent sites, respectively, thus
demonstrating the large contribution of the orbital magnetic moment. The discovery
of this unexpectedly large unquenched orbital moment on the Co site has been used
to explain the huge magnetocrystalline anisotropy of YCo5 . This particularly high

Table 12 Structural and magnetic properties of RFe5


Space
a(10−3 nm) c(10−3 nm) group TC (K) Ms (μB /f.u.) Meff (μB /f.u.) Ref.
ThFe5 511.6 404.6 P6/mmm 680 9.4 8.6 [273,
299]
NdFe5 494.6 417.0 P6/mmm 370 [264]
SmFe5 496.0 415.0 P6/mmm [300]
732

Table 13 Lattice parameters and magnetic properties of RCo5 compounds


Easy mag- Anisotropy
netization parame-
a(10−3 nm) c(10−3 nm) Spacegroup axis TSR (K) TC (K) Ms (μB /f.u.) ter(MJ/m3 ) Tcomp (K) Ref.
YCo5 493.7 397.8 P6/mmm c 977 8.33 4.2 K 4 K: K1 = [273,
8.01 (300 K) 7.4 MJ/m3 K2 = 296]
1.06 T −0.178 MJ/m3
300 K: K1 =
5.8 MJ/m3 K2 =
−0.358 MJ/m3
LaCo5 510.5 396.6 P6/mmm c 874 7.1 300 K: K1 = [322]
0.91 T 300 K 5.9 MJ/m3
CeCo5 492.2 401.6 P6/mmm c 794 5.5 300 K: K1 = [273,
300 K Ms = 5.8 MJ/m3 323,
0.85 T 324]
PrCo5 502.4 399.0 P6/mmm Cone 23◦ TSR =110 912 9.9 4 K: K1 = [273,
at 4.2 K to 9.0 at RT 1,12 T −7 MJ/m3 304,
c-axis 300 K: K1 = 325]
above TSR 8.1 MJ/m3
NdCo5 502.6 397.5 P6/mmm a for TSR1 =240 913 10.48 300 K: K1 = [301,
T<240 K, TSR2 =290 at RT Ms = 1,25 T 0.7 MJ/m3 323,
then c for 326]
T>290 K
SmCo5 497.4 398.5 P6/mmm c 984 8.11 4 K: K1 = [298]
7.56 (300 K) Ms = 26 MJ/m3
0.97 T 300 K: K1 =
18 MJ/m3
GdCo5 497.4 397.3 P6/mmm c 1008 1.57 300 K: K1 = [323,
2.32 (300 K) Ms = 4.6 MJ/m3 327]
0.19 T
J. P. Liu et al.
TbCo5 494.6(1) 397.9(1) P6/mmm a to cone to TSR1 = 365 980 0.51 110 22 [273,
c TSR2 = 450 1.76 (300 K) 322,
328]
DyCo5 492.6(2) 398.8(2) P6/mmm a to cone to TSR1 = 325 966 1.04 148 [323,
c TSR2 = 367 3.2 (300 K) 329]
HoCo5 488.1(3) 400.6(3) P6/mmm 4.2 K cone TSR = 200 1000 1.1 300 K: K1 = 85 [303,
84◦ to c 4.6 (300 K) MS = 3.6 MJ/m3 322,
above TSR 0.53 T 330]
ErCo5 487.0(4) 400.2(4) P6/mmm c – 986 0.4 300 K: K1 = – [273,
15 Metallic Magnetic Materials

5.6 at 300 K MS = 3.8 MJ/m3 322,


0.63 T 330]
TmCo5 486.3 401.7 P6/mmm c – 1020 1.9 – [273]
ThCo5 499.5 398.4 P6/mmm c – 415 7.35 – [264,
331,
332]
(1), (2), (3), and (4) compounds richer in Co having stoichiometry corresponding to R1−δ Co5+δ
733
734 J. P. Liu et al.

orbital moment on one Co site is attributed to its peculiar and asymmetric atomic
local environment.
In the RCo5 compounds, the Co sublattice anisotropy is dominant at high
temperature along the whole R series. For R = Pr, Nd, Tb, Dy, and Ho, a
spin reorientation occurs due to competition with the R anisotropy. The easy
magnetization direction is along the a-axis for R = Nd, Tb, and Dy, whereas for
Pr and Ho it is tilted away from the c-axis by 23◦ and 84◦ , respectively [303, 304].
Unlike for the RM5 compounds with M = Fe and Co, the RNi5 ones have
nonmagnetic M atoms, and the R elements are the only one carrying a magnetic
moment. For R = La, Y, Ce, Lu, and Th, the RNi5 phases are Pauli paramagnets,
whereas the PrNi5 exhibits Van Vleck paramagnetism. PrNi5 does not order
magnetically due to crystal electric field effects which lead to a nonmagnetic singlet
ground state lying below the first magnetic level [305]. Due to their exchange-
enhanced Pauli paramagnetism, YNi5 and CeNi5 have been considered as nearly
ferromagnetic [306, 307]. For the other R elements, magnetic order occurs at low
temperature, the largest ordering temperature being 32 K for GdNi5 [308]. Most of
the RNi5 compounds exhibit ferromagnetic order. It is worth noting that Ni atoms do
not carry a significant ordered magnetic moment in the RNi5 structure type, although
a small induced polarization of order 0.1 μB /Ni has been reported by several authors.
However, the Ni atoms are found to contribute to the paramagnetic state in particular
for the light R elements since the effective magnetic moments deduced from the
paramagnetic regime are much larger than for the free Pr3+ or Nd3+ ions [305, 309].
The main structural and magnetic properties of the other RNi5 -type compounds are
summarized in Table 14.
The experimentally determined second-order crystal electric field parameter of
the RNi5 compounds is compared in Table 15 to that of the RCo5 isotypes. It is
clearly seen that the negative sign is preserved when passing from one series to the
other; however, the values are much larger for the RNi5 compounds. A comparison
of the crystal electric field and magnetic behavior of the RNi5 and RCo5 compounds
has also been done on the basis of the Mössbauer spectroscopy at the R nucleus
(161 Dy, 166 Er, 169 Tm) [310]. For the RNi5 series of compounds, the set of the
crystal electric field parameters (up to the sixth order) has been determined from
investigation of single-crystal magnetization data. The corresponding values are
listed in Table 16 for R = Pr, Nd, Tb, Dy, Ho, Er, and Tm [311, 312]. For SmCo5 , a
A0 2 <r2 >/kB value of about −200 K has been determined from a polarized neutron
study on a single crystal [313]. The result is agreement with other investigations
[314–316]. Values of 0 K and 50 K were derived in SmCo5 for the higher-order
terms A0 4 <r4 >/kB and A0 6 <r6 >/kB , respectively [313].
The magnetocaloric properties of the RNi5 compounds have been systematically
studied (R = Pr, Nd, Gd, Tb, Dy, Ho, Er). Theoretical calculations are compared to
experimental data recorded on polycrystalline samples in the following reference
[317]. Around the ordering temperature, the magnetic entropy values of about

Smag = 7, 6, 7, 5 J/mol, K are obtained for R = Er, Ho, Dy, or Tb, respectively.
The adiabatic temperature change
Tad is 4, 6, 6 8, 9, 12 K for R = Nd, Gd,
Tb, Ho, Dy, and Er, respectively. An anomalous magnetocaloric effect is reported
Table 14 Lattice parameters and magnetic properties of RNi5 compounds
Easy
magnetization
a(10−3 nm) c(10−3 nm) Space Group axis TC (K) Ms (μB /f.u.) θp (K) Meff (μB /f.u.) Various Ref.
YNi5 489.1 396.1 P6/mmm – Pauli para. 0 −804 1.75 [306]
LaNi5 501.4 397.6 P6/mmm – Pauli para. 0 [307]
CeNi5 P6/mmm – Pauli para. 0 −275 1.34 [307]
PrNi5 495.7 397.6 P6/mmm – Van Vleck para. 0 9.2 [305]
495.2 397.6 P6/mmm 7 2.14 [309]
15 Metallic Magnetic Materials

NdNi5
SmNi5 492.4 397.4 P6/mmm c 27.5 0.7 [333]
EuNi5 491.1 397.5 P6/mmm – No order 0 [334]
GdNi5 490.6 396.8 P6/mmm c 32 6.90 31. 7.9 VZZ = 8.5 [335,
1021 nV/m2 336]
TbNi5 489.4 396.6 P6/mmm a 23 7.60 17.0 11.5 −
SM max = [337]
56 J/(mole
K)
H =
5T

Tad max =
7.3 K
DyNi5 487.2 396.8 P6/mmm b 11.6 8.5 8.2 [338]
HoNi5 487.2 396.6 P6/mmm 5 9 4.5 10.7 [311]
ErNi5 485.8 396.5 P6/mmm c 8 8.7 4.2 9.5 [339]
TmNi5 485.3 396.0 P6/mmm c 4.5 6.76 1.75 [340]
YbNi5 484.7 395.7 P6/mmm a 0.55 3.9 [341]
LuNi5 481.6 398.2 P6/mmm – Pauli para. – [342]
ThNi5 497.0 401.0 P6/mmm – Pauli para. 0 [332, 343,
344]
735
736

Table 15 Comparison of some second-order crystal electric field parameters reported for RCo5 and RNi5 compounds
R Pr Nd Sm Gd Tb Dy Er Tm
RNi5 A0 2 <r2 >kB −270 −521 [312] −520 [333] – −501 [312] −499 [345] −271 [312] −376 [312]
(K) [312]
RCo5 A0 2 <r2 >kB – −455 [345] −213 [313, 314, −206 [310] −124 [315] −167 [315] – –
(K) 346]
Note that the A0 2 parameter has been found to decrease upon increasing the Co stoichiometry [310]
J. P. Liu et al.
15 Metallic Magnetic Materials 737

Table 16 Crystal electric field parameters and exchange interaction coefficients reported for some
RNi5 and RCo5 single-crystal compounds at 4.2 K
nR-M (T/μB ) B0 2 (K) B0 4 (10−2 K) B0 6 (10−4 K) B6 6 (10−4 K)
NdCo5 [301, 347] 3.71 0.25 −5.0 −124
PrCo5 [348] 1.21 2.4 10
HoCo5 [349] 0.69 0.14 0.04 –
PrNi5 [305] 2.1 5.84 4.53 8.86 314
NdNi5 [309] 3.35 1.45 −3.50 135
TbNi5 [311] 3.84 −0.039 −0.40 −4.0
DyNi5 [312] 2.30 0.22 0.10 2.7
HoNi5 [311] 1.15 0.19 −0.02 −3
ErNi5 [311] −0.69 0.0 0.53 3.5
h [339] −3.80 −1.26 1.83 −7.6

for PrNi5 [317]. Magnetocaloric effect has also been investigated in a HoCo5
single crystal in the temperature region around the spin reorientation. A significant
anisotropic behavior has been reported when applying the magnetic field along
different directions [318].
The RCo5 compounds have been widely investigated by neutron diffraction
experiments, and results are given in Table 17 for the compounds with R = Y, La,
Ce, Nd, Sm, Tb, Ho, and Th. ThCo5 has been found to present a sharp itinerant
electron metamagnetic behavior in its the magnetization curve, a feature highly
sensitive to the Co stoichiometry [319]. At 4.2 K, whereas in ThCo5 the critical
field occurs at 2.2 T or ThCo5.1 , the Co magnetic moments are already in a high
magnetization state. The Co magnetic moment increases from about 1.0 μB to
1.5 μB /atom at the transition [320]. The effect of pressure on the metamagnetic
transition has also been reported to be large [321], the transition being shifted to
higher critical field at a rate of 2.8 T/kbar.
The R2 M17 (M = Fe, Co, Ni) compounds have the richest M content of the
binary phase diagram and crystallize in superstructures deriving from the RM5
(CaCu5 ) type by an ordered substitution of two Co atoms (forming a dumbbell pair)
for R atoms following the sequence 3 RM5 – R + 2M => R2 M17 . Depending on
the relative size of the R and M atoms, these phases may exist in either hexagonal
Th2 Ni17 type (P63 /mmc) [343] for M = Fe and Ni or rhombohedral Th2 Zn17
type (R3m) structure [351] for M = Fe, Co. Some compounds can be obtained
in either allotropic form depending on the synthesis conditions or composition.
Whereas the Th2 Zn17 type is a well-ordered structure, the Th2 Ni17 type can present
some disorder and is better described as the R2−δ M17+2δ [352–355] as a result
of the replacement of part of the R atoms by a dumbbell M pair. Magnetic data
concerning the R2 M17 compounds are listed in Tables 18 and 19 for M = Fe and
Co, respectively. Partial replacement of these three transition metal elements is
possible in particular by Mn. The properties of such pseudo-binary compounds are
not described below due to space limitation; for more details, the reader is referred,
for example, to [356].
738

Table 17 Magnetic structure determined from neutron diffraction for RCo5 compounds
Th[319] Th[319] Th[320] Th[320]
RCo5 Y[350] Y[302] La[295] Ce[328] Nd[301] Sm[313] Sm[313] Tb[328] Tb[328] Tb[328] Ho[349] Ho[349] ThCo5 ThCo5 ThCo5 ThCo5.10
T (K) 4.2 300 300 4.2 4.2 300 4.2 300 453 4.2 300 110 4.2
@
4.4 T
Easy c c c Perp to c c a a c Perp c c c c c
direc- c to c
tion
MR −0.4 – <0.3 2.82 3.8 0.04 8.35 6.30 4.10 10.0 4.8 – – – –
(μB )
MCo 1.68 1.77 1.60 1.3 1.88 1.86 1.86 1.55 1.35 1.45 1.65 1.7 1.21 1.19 1.4 1.63
2c
(μB )
MCo 1.67 1.72 1.76 1.3 1.91 1.75 1.75 1.70 1.55 1.45 1.65 1.7 0.96 1.06 1.4 1.63
3g
(μB )
J. P. Liu et al.
Table 18 Lattice parameters and magnetic properties of R2 Fe17 compounds (EMD stands for easy magnetization direction)
a −ΔSmax
(10−3 c(10−3 V(10−3 Type/space TC Ms dTc /dP d T /dP K1 (4 K) K2s (4 K) (J/kg
nm) nm) nm) group (K) EMD (μB /f.u.)4 K (K/kbar) (K/kbar) (MJ/m3 ) (MJ/m3 ) K) Ref.
Th2 Fe17 857.1 1247.4 798.2 Th2 Zn17 /R3m 315 Ferro 30.5 −4 [386,
387]
Ce2 Fe17 848.9 1241.3 774.8 Th2 Zn17 /R3m 225 Plane 29.7 −2.5 −40 [358,
(Tc ) fan to 359,
15 Metallic Magnetic Materials

heli- 361,
cal 362,
388]
Pr2 Fe17 858.1 1246.3 794.8 Th2 Zn17 /R3m 293 a 36.4 −3 −3.7 0.55 [372,
387]
Nd2 Fe17 857.9 1246.1 794.3 Th2 Zn17 /R3m 326 38.0 −3.6 KNd 1 = −55 KNd 2 = 3 [370,
23 (@2 T) 389]
Sm2 Fe17 855.4 1244.3 788.4 Th2 Zn17 /R3m 385 a 35.4 −3 <0.1 1.72 [372,
(@1.5 T) 390]
Gd2 Fe17 853.9 1243.3 785.5 Th2 Zn17 /R3m 461 a-b 21.5 −2 0.5 0.89 [390,
(Tc ) plane (@1.5 T) 391]
Tb2 Fe17 846.2 831.5 515.6 Th2 Ni17 /P63 /mmc 425 b 20.0 0.3 −5 1.32 [390,
(Tc ) (@1.5 T) 392,
393]
Dy2 Fe17 846.2 830.0 514.6 Th2 Ni17 387 a 16.2 −4.8 −K1 −2K2 −3K3 1.59 [390,
P63 /mmc (Tc ) = −1.5 (@1.5 T) 394,
(80 K) 395]
Ho2 Fe17 845.3 828.9 513.1 Th2 Ni17 335 b 19.5 −5.6 1.75 [371,
P63 /mmc (Tc ) −K1 −2K2 −3K3 394]
= −2 (80 K)
739

(continued)
740

Table 18 (continued)
a −ΔSmax
(10−3 c(10−3 V(10−3 Type/space TC Ms dTc /dP d T /dP K1 (4 K) K2 (4 K) (J/kg
nm) nm) nm) group (K) EMD (μB /f.u.)4 K (K/kbar) (K/kbar) (MJ/m3 ) (MJ/m3 ) K) Ref.
Er2 Fe17 843.8 828.2 510.7 Th2 Ni17 300 a 17.9 −4.3 −K1 −2K2 −3K3 1.68 [390,
P63 /mmc (Tc ) = −1.5 (80 K) (@1.5 T) 393,
394]
Tm2 Fe17 842.0 827.2 507.9 Th2 Ni17 274 c 21.2 6.6 −2.2 [368,
P63 /mmc (Tc ) T<74 K 396,
(TSR ) a 397]
T> TSR
Yb2 Fe17 841.4 824.9 507.2 Th2 Ni17 278 b [398]
P63 /mmc (Tc )
Lu2 Fe17 840.6 827.5 506.4 Th2 Ni17 270 b fan 34.2 −1.9 −48 −2.3 4.1 [365,
P63 /mmc (Tc ) LT to (@5 T) 394,
helical 1.5 399]
(@2 T)
Y2 Fe17 842.2 830.6 515.1 Th2 Ni17 327 Ferro 34.6 −4.7 −2.3 −0.11 2.4 [355,
R3m also or fan (@2 T) 364,
found as 4 to 5 369,
Th2 Zn17 (@5 T) 399]
P63 /mmc
J. P. Liu et al.
Table 19 Lattice parameters and magnetic properties of R2 Co17 compounds
K1 K2 K1 K2
(MJ/m3 ) (MJ/m3 ) (MJ/m3 ) (MJ/m3 )
15 Metallic Magnetic Materials

a (10−12 m) c (10−12 m) Space Group TC (K) Ms (μB /f.u.) BA (T) 300 K 300 K 77 K 77 K Ref.
Th2 Co17 843.8 1225.4 R3m 1035 24.14 [264]
Ce2 Co17 836.8/837.1 1220.4/813.6 R3m/P63 /mmc 1083 26.1 −0.6 [351,
1.15 T at 378,
300 K 400]
Pr2 Co17 843.8/ 1225.3 R3m 1171 31.0 −0.6 [351,
1.38 T at 378,
300 K 400]
Nd2 Co17 842.2 1224.6 R3m/P63 /mmc 1150 30.5 −1.1 [351,
1.39 T 378,
400]
Sm2 Co17 838.5/826.0 1221.4/851.5 R3m/P63 /mmc 1190 20.1 3.3 0.35 [351,
1.2 T at 378,
300 K 400]
Gd2 Co17 837.7/8.351 1219.8/8.125 R3m/P63 /mmc 1209 14.4 −0.5 −0.47 0.05 [351,
0.73 T 378,
300 K 400]
Tb2 Co17 835.7/834.8 1218.6/812.5 R3m/P63 /mmc 1180 (Tc) 10.7 −3.3 [351,
0.68 T 378,
300 K 400]
741

(Continued)
742

Table 19 (Continued)
K1 K2 K1 K2
(MJ/m3 ) (MJ/m3 ) (MJ/m3 ) (MJ/m3 )
a (10−12 m) c (10−12 m) Space Group TC (K) Ms (μB /f.u.) BA (T) 300 K 300 K 77 K 77 K Ref.
Dy2 Co17 834.6/832.8 1281.0/812.5 R3m/P63 /mmc 1152 (Tc) 8.3 −2.6 −2.8 0.2 [351,
0.70 T at 378,
300 K 400]
Ho2 Co17 833.2/832.0 1220.1/811.3 R3m/P63 /mmc 1173 (Tc) 7.7 −1.0 −2.0 0.15 [351,
0.83 T at 378,
300 K 401]
Er2 Co17 –/831.0 –/811.3 P63 /mmc 1187 (Tc) 10.1 18 T 0.41 1.7 0.4 [351,
0.90T at (295 K) 378,
300 K 402]
Tm2 Co17 –/828.5 –/809.5 P63 /mmc 1182 (Tc) 11.3 0.5 2.7 0.4 [378,
1.13 T at 401]
300 K
Yb2 Co17 –/830.9 –/809.6 P63 /mmc 1180 (Tc) 20.46 9.9 T 0.38 0 [398,
(4.2 K) (295 K) 402,
26.8 403]
(295 K)
Lu2 Co17 –/824.7 –/809.3 P63 /mmc 1175 (Tc) 26.9 −0.20 0.02 −0.58 0.09 [384,
1.87 T 401]
300 K
Y2 Co17 835.5/834.1 1218.3/812.5 R3m/P63 /mmc 1167 27.8 −0.3 <0.01 −0.58 0.05 [400]
1.25 T
300 K
J. P. Liu et al.
15 Metallic Magnetic Materials 743

The R2 Fe17 compounds have attracted much interest due to their large mag-
netization mostly resulting from the large Fe content. However, their relatively
low ordering temperature ranging from 225 (Ce) to 461 K (Gd) hampers their
use as magnets at room temperature. The low Curie temperatures and unfavorable
magnetic anisotropy can be resolved by introducing N or C in interstitial sites
surrounding the rare-earth atoms. Generally, the R2 Fe17 compounds are ferromag-
netic for light rare-earth and ferromagnetic for heavy rare earth. Exceptions are
Ce2 Fe17 and Lu2 Fe17 which have been reported to be at the limit of ferro- and
antiferromagnetic ordering. In a limited temperature range, noncollinear magnetic
structures were observed in Ce2 Fe17 , Lu2 Fe17 , or even Tm2 Fe17 compounds where
negative exchange interactions are due to short Fe-Fe distances [357]. At low
temperature, these compounds can present helimagnetic structures, the Fe magnetic
moment lying in the basal (a,b) plane. The orientation of the magnetization of the
ferromagnetic planes rotates from one (001) plane to the following one forming
a helix whose period varies with temperature [357–359]. An antiferromagnetic
behavior occurs for Lu2 Fe17 between the magnetic ordering temperature 270 K and
the second critical temperature Θ T of 140 K below which it becomes ferromagnetic
[357]. This second critical temperature decreases rapidly upon application of
pressure dΘ T /dp = −48 K/kbar, whereas the ordering temperature is much less
sensitive to the pressure [360] dTN /dp = −1.9 K/kbar. Recent studies on Ce2 Fe17
[361, 362] established that below 118 K additional superstructure reflexions appear
in the antiferromagnetic state leading to a doubling of the chemical unit cell along
the c direction in hexagonal notation with the same space group R3m. dΘ T /dp
= −40 K/kbar.
Y2 Fe17 is a collinear ferromagnetic with magnetic moments lying in-plane up
to the TC ; however, application of high pressure induces a change of its magnetic
structure to helical incommensurate phases [363]. A similar effect of pressure has
been reported for Ce2 Fe17 and Lu2 Fe17 [359, 364, 365]. A helimagnetic ground
state has been induced in Y2 Fe17 and Lu2 Fe17 intermetallics by hydrostatic pressure
above 1.05 and 0.3 GPa, respectively [364].
The TC of R2 Fe17 ferromagnetic compounds decreases upon increasing applied
pressure, and values of dTc /dP = −3.6 and −4.7 K/kbar have been reported
for Nd2 Fe17 and Y2 Fe17 , respectively. The iron sublattice magnetization is about
34 μB /formula unit for both R = Y and Lu compounds, corresponding to a
mean Fe moment of 2 μB in the ordered state. In the paramagnetic state, an
effective magnetic moment of 3.87 μB /Fe atom has been reported indicating the
itinerant electron character of the magnetism in Y2 Fe17 [366]. All the R2 Fe17
compounds present planar anisotropy, except Tm2 Fe17 . A spin reorientation occurs
at 74 K upon heating, temperature above which the magnetic moments tilt toward
the (a,b) plane [357, 367, 368]. Magnetocrystalline anisotropy has been deter-
mined on single-crystal Y2 Fe17 [369] and Nd2 Fe17 [370] leading to values of
KFe 1 = 0.255 MJ/m3 and KFe 2 = −3.11 MJ/m3 for the Fe sublattice anisotropy
against KNd 1 = −57 MJ/m3 and KNd 2 = 26 MJ/m3 . The magnetocrystalline
anisotropy of Ho2 Fe17 has been determined at 4.2 K: K1 = −5.6 MJ/m3 and K2 =
1.7 MJ/m3 [371]. For Sm2 Fe17 , a value of K1 = −3 MJ/m3 has been estimated from
high field measurements on powder sample [372]. FOMP (first-order magnetization
744 J. P. Liu et al.

processes) have been reported to occur at low temperature for several R2 Fe17
compounds [373]. Neutron diffraction experiments and Mössbauer spectroscopy
analysis agree to show that the largest Fe magnetic moment is observed on the
dumbbell Fe sites (6c and 4e in the R3m and P63 /mmc, respectively), the other
Fe moment being close to 2 μB in the R2 Fe17 compounds [374, 375]. The
magnetocaloric properties of several R2 Fe17 compounds have been determined; the
values are also listed in Table 18.
In the R2 Co17 series, the compounds show ordering at high temperature ranging
from 1035 K for Ce to 1209 K for Gd, the Co magnetic moments being coupled
ferromagnetically for R = Lu, Y, Ce, or Th. According to room temperature neutron
diffraction experiments on Y2 Co17 , the Co magnetic moment is about 1.88 μB /Co
for the different crystal sites and even 2.1 μB on the 6c position [376]. Y2 Co17
presents a planar anisotropy featured by K1 = −0.34 MJ/m3 at room temperature
and a saturation magnetization of 1.25 T. Thanks to its high magnetization of 1.2 T
and large magnetic anisotropy K1 = 3.3 MJ/m3 at room temperature, Sm2 Co17
exhibits exceptional magnetic properties for high-temperature, high-performance
permanent magnet applications [377–379] leading to a maximum energy product
of 287 kJ m−3 [377, 379]. Many efforts have been made to play with substitutions
and/or additions of elements such as Zr, Cu, and Fe, in order to improve the
synthesis process and get optimized microstructure leading to high-performance
magnets [380].
Compounds containing R = Tm, Er, and Yb are the only other ones to exhibit
positive values of the first-order anisotropy constant at 4.2 K K1 = 0.189, 0.5 MJ/m3 ,
and 0.34 MJ/m3 , respectively [381]. The anisotropy parameters determined at either
300 K or 77 K are listed in Table 19 together with the ordering temperature
and other magnetic properties of the R2 Co17 compounds [273]. At 77 K, the
anisotropy parameter of the Co sublattice estimated from the compounds with R
= Y and Lu is K1 = −0.58 MJ/m3 , and values of K2 = 0.05 MJ/m3 and K2 =
0.09 MJ/m3 are reported for Y and Lu, respectively [382, 383]. In a recent study on
a Lu2 Co17 single crystal, a spontaneous magnetic moment of 27.6 μB /formula unit
corresponding to about 1.62 μB /f.u. per Co atom has been reported at 5 K [384], and
the easy plane magnetocrystalline anisotropy is confirmed as the ground state since
K1 = −0.57 MJ/m3 and K2 = −0.058 MJ/m3 . As a consequence of this second-
order phase transition, the compound becomes easy axis at elevated temperature
between 730 K and the Curie temperature [384]. A polarized neutron diffraction
study performed at 5 K on single crystal of Er2 Co17 leads to the following magnetic
moments values: mEr (2b) = 8.3 μB , mEr (2d) = 9.0 μB , mCo (4f ) = 2.1 μB , mCo
(6g) = 1.6 μB , mCo (12j) = 1.9 μB , and mCo (12k) = 1.2 μB [385].
R2 M14 B compounds are found for M = Fe, Co and X = B or C in a tetragonal
structure whose prototype is Nd2 Fe14 B crystallizing in the P42 /mnm space group.
The crystal structure has two inequivalent sites for R elements, six for the iron
and one for the boron atoms [404–406]. These phases are line compounds with
a defined stoichiometry. The lattice parameters and the main magnetic properties
corresponding to the R2 Fe14 B, R2 Co14 B, and R2 Fe14 C compounds are listed in
Tables 20, 21, and 22, respectively.
Table 20 Magnetic properties and room temperature lattice parameters of R2 Fe14 B-type compounds
EMD Ms K1 (MJ/
a(10−3 c(10−3 TC at TSR Ms (μB /f.u.) μ0 Ms (T) K1 K2 4.2 K (μB /f.u.) μ0 Ms (T) m3 )
nm) nm) (K) 4K (K) 4.2 K 4.2 K BA (T)4.2 K 4.2 K(MJ/m3 (MJ/m3 ) 300 K 300 K 300 K Ref.
15 Metallic Magnetic Materials

Th2 880.5 1218.7 487 c 29.8 1.4 1.0 24.7 1.25 1.3 [422]
Fe14 B
La2 882 1234.0 542 c 30.8 1.50 1.3 0.7 28.4 1.39 1.1 [423]
Fe14 B
Ce2 876 1211.0 427 c 30.0 1.52 3.0 1.8 23.9 1.17 1.8 [423]
Fe14 B
Pr2 880 1223.0 565 c 37.6 1.76 29 55 −87 31.9 1.55 5.0 [424]
Fe14 B
Nd2 Fe14 B 880 1220.0 592 Canted 135 37.5 1.86 35 −16.6 39.3 32.5 1.61 4.9 [412,
30◦ /c 425]
Sm2 Fe14 B 880 1215.0 621 a 33.3 1.57 40 −26.0 30.2 1.51 −12.0 [424,
426,
427]
Gd2 Fe14 B 879 1209.0 664 c 17.9 0.9 2.1 0.7 17.5 0.83 1.0 [428]
Tb2 Fe14 B 877 1205.0 620 c 13.2 0.65 40 6.9 14.0 0.92 3.4 [423]
Dy2 Fe14 B 876 1201.0 589 c 11.3 0.58 23 3.8 14.0 0.72 4.2 [423,
424]
Ho2 Fe14 B 876 1199.0 569 Canted 57 11.2 0.57 18 −1.1 4.4 15.9 0.80 2.0 [417,
24◦ /c 423]
745

(continued)
746

Table 20 (continued)
EMD Ms K1 (MJ/
a(10−3 c(10−3 TC at TSR Ms (μB /f.u.) μ0 Ms (T) K1 K2 4.2 K (μB /f.u.) μ0 Ms (T) m3 )
nm) nm) (K) 4K (K) 4.2 K 4.2 K BA (T)4.2 K 4.2 K(MJ/m3 (MJ/m3 ) 300 K 300 K 300 K Ref.
Er2 Fe14 B 873 1195.0 554 a 328 13.1 0.66 16 −1.4 17.7 0.90 1.3 [417,
423]
Tm2 Fe14 B 873 1193.0 545 a 314 18.2 0.93 17 −3.6 22.6 1.15 – [423,
429]
Yb2 Fe14 B 871 1185.0 527 a 120 ∼23 ∼1.2 ∼23 – [429–
431]
Lu2 Fe14 B 876 1200.0 534 c 28.5 1.47 2.4 1.2 22.5 1.13 1.5 [423,
432]
Y2 Fe14 B 876 1200.0 566 c 29.5 1.55 1.5 0.8 26.05 1.44 1.1 [428]
TSR Temperature of spin reorientation transition
J. P. Liu et al.
Table 21 Magnetic properties and room temperature lattice parameters of R2 Co14 B-type compounds
Ms
a(10−3 c(10−3 TC TSR μ0 Ms (T) K1 (MJ/m3 ) (μB /fu) K1 (MJ/ K2
nm) nm) (K) (K) 300 K 300 K EMD 4.2 K μ0 Ms (T)4.2 K BA (T)4.2 K m3 )4.2 K (MJ/m3 ) Ref.
La2 Co14 B 867 1201 955 1.00 −1.4 (001) 20.4 1.1 3.4 −1.4 [427]
plane
15 Metallic Magnetic Materials

Pr2 Co14 B 863 1187 995 664 1.09 6.0 c to (001) 24.8 1.3 75 [433]
plane
Nd2 Co14 B 864 1186 1007 37 1.06 2.2 Canted 25.7 1.3 30 −1.4 19 [427,
543 15◦ to c 434,
(4.2 K) 435]
then c
(300 K) to
perp. to c
Sm2 Co14 B 861 1179 1029 (001) 18.1 1.0 [436]
plane
Gd2 Co14 B 861 1176 1050 −1.3 (001) 5.4 0.3 14 −1.4 [427,
plane 436]
Tb2 Co14 B 860 1173 1035 795 c to (001) 0.1 [436]
plane
Dy2 Co14 B 876 1201 [433]
Y2 Co14 B 860 1171 1015 1.00 −1.4 (001) 19.2 1.0 3.0 −1.3 [410,
plane 436]
747
748

Table 22 Magnetic properties and room temperature lattice parameters of R2 Fe14 C-type compounds
a(10−3 c(10−3 Ms μ0 Ha μ 0 Ms μ0 Ha
nm) nm) TC (K) EMD TSR (K) (μB /f.u.)4 K (T)4.2 K (T)300 K (T)300 K Ref.
La2 Fe14 C 882 1214
Ce2 Fe14 C 874 1184 345 c 23.9 [437]
Pr2 Fe14 C 882 1204 513 c 33.7 1.27 14.8 [438]
Nd2 Fe14 C 880.9 1205.0 530 Cone to 140 31.4 1.41 10.1 [438]
c-axis
Sm2 Fe14 C 879.8 1194.5 580 (a,b) plane 30.2 1.41 8.5 [432]
Gd2 Fe14 C 879.1 1189.3 630 c 18.1 2 0.73 3.38 [432]
Tb2 Fe14 C 877.0 1186.5 585 c 12.0 40 0.58 19.4 [439]
Dy2 Fe14 C 875.4 1182.6 555 c 10.5 0.62 15.4 [432]
Ho2 Fe14 C 873.9 1179.7 525 Cone to 35 10.9 0.0.73 8.2 [432, 440]
c-axis
Er2 Fe14 C 873.0 1177.5 510 Plane to 325 12.5 0.95 [441, 442]
c-axis
Tm2 Fe14 C 872.1 1174.9 500 [100], then 308 18.4 [443]
[001] for
T>TSR
Lu2 Fe14 C 871.3 1172.2 495 c 27.2 2.5 1.16 3.25 [432, 444]
Y2 Fe14 C 876 1186 520 c [445]
J. P. Liu et al.
15 Metallic Magnetic Materials 749

The R2 Fe14 B compounds have density ranging from 7.4 to 8.86 g/cm3 for
R = La and Th, respectively. They have attracted much interest due to their large
magnetization mostly resulting from their large Fe moments and the discovery that
due to the large magnetocrystalline anisotropy, one can obtain large coercive field in
alloys containing the Nd2 Fe14 B-type phase. At room temperature, a large saturation
magnetization of 1.61 T is reached for Nd2 Fe14 B, which is an interesting property
since the energy product of a magnet scale with Ms 2 [407]. The anisotropy constant
K1 of Nd2 Fe14 B is about 4.9 MJ/m3 at room temperature [408–410], whereas the
best magnets present excellent performances such as energy products larger than
400 kJ/m3 . Their ordering temperature, ranging from 427 K (Ce) to 664 K (Gd), is
high enough to permit their use as permanent magnets at room temperature [408,
409]; however, these Curie temperatures are relatively low compared with those
of other ferromagnetic materials used in permanent magnets (e.g., the R-Co-type
magnets). Generally, the R2 Fe14 B compounds are ferromagnetic for light rare-earth
elements and ferrimagnetic for heavy rare-earth elements. The compounds with R =
La, Ce, Lu, Y, and Th are ferromagnetic, their magnetic properties being determined
by the Fe sublattice only. They exhibit magnetocrystalline anisotropy of about 2 T
[410–412] at room temperature and MS of typically 30 μB /f.u. (1.5T) at 4.2 K and
28 μB /f.u. (1.4T) at 300 K (Table 20). This corresponds to a large value of the mean
Fe moment of about 2 μB at room temperature and even 2.14 μB at 4.2 K. The
Mössbauer effect spectroscopy [413], neutron diffraction experiments [374, 414–
419], and theoretical calculations [420, 421] agree to reveal that the magnitude of
the Fe magnetic moment varies significantly for the inequivalent atomic positions,
the largest value (up to 2.8 or even about 3.0 μB ) being reported for the 8j2 position
[374, 417, 420]. At 4.2 K, the Fe sublattice K1 anisotropy parameter ranges from 0.8
to 1.2 MJ/m3 for the Fe sublattice depending on the R element (Table 20). Sm2 Fe14 B
is the only compound exhibiting a planar character, its easy magnetization direction
remaining within the (a,b) plane up to TC [448]. For the other R elements, the Fe
sublattice anisotropy dominates at high temperature favoring uniaxial anisotropy.
More complex behavior is observed at lower temperature due to the increasing
influence of the R sublattice anisotropy that is reinforced at cryogenic temperatures.
The magnetic phase diagram of the R2 Fe14 B compounds is schematically repre-
sented in Fig. 11 [446]. Typical thermal variation of the K1 anisotropy parameter
is shown in Fig. 12 for R = Nd, Pr, and Dy. At low temperature, several R2 Fe14 B
compounds undergo a spin reorientation phenomenon occurring at 135 K, 57 K,
328 K, 314 K, and 120 K for R = Nd, Ho, Er, Tm, and Yb, respectively [408, 409].
For R = Er, Tm, and Yb, the spin reorientation results from the competition between
the R and Fe sublattice magnetocrystalline anisotropy, the R one being dominant
at low temperature inducing an easy plane magnetization at 4.2 K as witnessed
by the anisotropy parameters reported in Table 20. For R = Nd and Ho, the spin
reorientation is of different origin; the magnetic moments are tilted away from the
c-axis below TSR [416, 417]. This phenomenon is a consequence of the different
temperature variation of the various magnetocrystalline anisotropy terms associated
750 J. P. Liu et al.

Fig. 11 Schematic representation of the magnetic phase diagram of the R2 Fe14 B [446, 447]

to the R sublattice [449, 450]. In Nd2 Fe14 B, this results in a non-strictly collinear
magnetic structure and different tilt angles within the (1–10) plane of the Fe and
Nd sublattices. Due to their nonequivalent atomic positions and the consequently
different crystal electric field parameters, the Nd sites can have different tilt angles
too [417] as shown in Fig. 13. The existence of this spin reorientation for R = Nd is
a drawback for low temperature applications as permanent magnet. Consequently,
Pr2 Fe14 B is often preferred due to its large uniaxial anisotropy even at 4.2 K. It is
worth to point out that neutron diffraction revealed that due to the spin reorientation,
a lowering of symmetry is observed [417, 418]. As a consequence, the R 4f and 4g
sites are further subdivided magnetically into f1 , f2 , g1 , and g2 sites as derived from
inelastic neutron scattering measurements [451].
The R6 M13 A phases (A= Cu, Ag, Au, Si, Ge, Ga, Al, Sn, Sb, Tl, Pb, Bi)
have been discovered in the frame of the search of phase stability; they can be
observed as secondary phase in some starting materials for R2 Fe14 B-type permanent
magnets when some A additives are used [453, 454]. Often referred to as δ phase,
the R6 M13 A compounds attracted much interest since such secondary phases have a
critical influence on the magnetic performances of the magnets (i.e., the coercitivity)
[455, 456]. These R6 M13 A-type compounds crystallize in the La6 Co13 Ga-type
structure [453, 454, 457] an ordered version of the Nd6 Fe11 Ga3 and are observed
15 Metallic Magnetic Materials 751

40

Pr2Fe14B
20
K1 (MJ/m3) Dy2Fe14B

Nd2Fe14B
-20

-40
0 200 400 600
Temperature (K)

Fig. 12 Temperature dependence of the anisotropy constant of the R2 Fe14 B with R=Pr, Nd, and
Dy [452]

Fig. 13 Magnetic structure of the R2 Fe14 B R = Ho (left), Nd (center), and Er (right) as


determined from single-crystal neutron diffraction at 20 K [417]

for light R elements such as La, Pr, Nd, and Sm. Such tetragonal structure retains
the I4/mcm space group and is shown in Fig. 14. The R elements are located on
two positions 16l and 8f, whereas the iron atoms are found on four inequivalent
sites 4d, 16k, 16l1 , and 16l2 . Finally, the A atom sits on the 4a position. Such
structure can be seen as a multilayer structure where slabs of iron alternate with
slabs of rare-earth atoms. The lattice parameters are listed in Table 23 together with
some magnetic properties of the R6 Fe13–x Ax -type compounds. Depending on the A
nature, this crystal structure can accommodate a certain homogeneity range leading
to R6 Fe13–x Ax formula. For example, with A = Ga, it has been shown that Ga is
filling up successively the 4a Si and the 16l Fe4 sites of the prototype Nd6 Fe13 Si
752 J. P. Liu et al.

Fig. 14 Schematic representation of the crystal structure and collinear antiferromagnetic struc-
tures in Pr6 Fe13 Sn (left part) and Nd6 Fe13 Sn (right part) with the moments, respectively,
perpendicular to or along the c-axis. The inequivalent atomic positions are indicated with different
symbols for R, Fe, and X = Sn [464]

structure with increasing x [458]. Even larger solubility limit has been reported for A
= Al since Nd6 Fe9.5 Al4.5 compound has been reported [459, 460]. On the contrary
for A = Cu, a line compound with x = 1 [461] is found. It is worth to point out that
low amount of heavier rare-earth elements like Gd or Dy can be partially substituted
to the light one in the R6 Fe13–x Ax (A=Ga, Al) [461].
The R6 M13 A compounds exhibit antiferromagnetic properties, properties that
have sometimes been hidden by the presence of traces of ferromagnetic-type
impurities like R2 Fe17 [454]. For A = Au, Ag, Cu, Si, and Ga, the ordering
temperature is around 415 K [461]. The ordering temperature decreases upon
increasing the A content: in R6 Fe13–x Gax compounds, TN = 433 and 365 K for
x = 1 and 2, respectively. Similar trend has been reported for other systems since
15 Metallic Magnetic Materials 753

Table 23 Lattice parameters and magnetic ordering of R6 Fe13–x Ax compounds


R6 Fe13–x Ax a (10−3 nm) c (10−3 nm) TN (K) TSR (K) Ms (μB /f.u.)4 K Ref.
La6 Fe11 Al3 822.0 2382.0 230 [461]
Ce6 Fe11 Al3 819.0 2310.1 230 [466]
Pr6 Fe13 Ag 812.1 2282.0 380 [462]
Pr6 Fe13 Au 809.6 2267.2 380 [462]
Pr6 Fe13 Si 805.4 2287.3 [470]
Pr6 Fe13 Ge 806.6 2296.8 303 [466, 471]
Pr6 Fe13 Ga 809.9 2308.4 [458]
Pr6 Fe12 Ga2 809.9 2308.4 [458]
Pr6 Fe13 Sn 809.8 2341.8 393 [464]
Pr6 Fe13 Bi 811.6 2351.6 453 [472]
Nd6 Fe13 Cu 811.1 2230 419 38.2 [461]
Nd6 Fe13 Ag 808.5 2259.1 415 41.2 [473]
Nd6 Fe13 Au 808.4 2260 411 42.5 [461, 470]
Nd6 Fe13 Si 802.9 2270.7 421 100 >32 [461, 463]
Nd6 Fe13 Ga 807.2 2295 433 41.9 [458, 461]
Nd6 Fe12 Ga2 809.2 2298 365 [458]
Nd6 Fe13 Sn 808.8 2329.8 433 [464]
Sm6 Fe13 Bi 805.7 2331.6 [472]

for R = Pr TN = 390, 335, and 320 K for x = 1, 2, and 3, respectively ([458] and
reference therein).
Several neutron diffraction experiments have reported the magnetic structure
of some R6 M13 A-type compounds. Pr6 Fe13 Ag and Pr6 Fe13 Au exhibit collinear
antiferromagnetic structure [462], the magnetic moments being confined within the
(a,b) plane but change sign collectively when passing to the next (R-Fe) block
along the c direction, that is, crossing the A layer. The magnetic moments of the
four iron and two Nd crystallographic sites are ferromagnetically coupled within
one block along the c-axis, and the resulting magnetic moments of this block
are antiferromagnetically coupled with that of the adjacent block along the c-axis
through the A atom layer. Antiferromagnetism is preserved over the anticentering
magnetic lattice translation so that two successive blocks have opposite signs.
The average Fe magnetic moment is about 2.2 μB in agreement with Mössbauer
spectroscopy results [463]. Similar magnetic structures have been reported for
Pr6 Fe13 Sn [464] (see Fig. 14), whereas the Nd isotype compound presents similar
alternating magnetic blocks but with an alignment of the magnetic moments along c.
The prototype Nd6 Fe13 Si compound presents an original behavior with a spin
reorientation occurring at about 100 K. Above and below 100 K, the magnetic
moments are found to be parallel to the c-axis and within the (a,b) plane of
the tetragonal unit cell, respectively. This reorientation occurs via a first-order
spontaneous zero field spin reorientation and is accompanied with an anisotropic
unit cell expansion at the spin reorientation [463]. A detailed 57 Fe Mössbauer
spectroscopy analysis of Nd6 Fe13 Si compound is also described in the same article.
754 J. P. Liu et al.

The lattice parameters and main magnetic properties of the R6Fe13-xAx compounds
are summarized in Table 24. For Nd6 Fe13–x Gax (x = 0 to 2) compounds, an easy-
axis antiferromagnetic structure is found below the Néel temperature and down
to a spin reorientation temperature below which an easy cone antiferromagnetic
structure is observed with a tilt angle of 9.2◦ versus c-axis at 1.5 K [458]. This
antiferromagnetic coupling between the magnetic blocks on both sides of the A
layer can be broken by the application of an external magnetic field, thus leading
to a metamagnetic transition occurring via a first-order magnetic phase transition
[465–467]. The magnetization obtained in such a forced ferromagnetic state has
been reported to be over 41 μB /f.u. for R6 Fe13 A compounds [461]. A large
hysteresis in the field dependence of the magnetization is present in all compounds
including La6 Fe11 Al3 , indicating the role of the Fe-sublattice anisotropy [461]. The
influence of Co for Fe substitution on magnetic properties and thermal expansion
of Nd6 Fe13 Si intermetallic compounds has been reported [468, 469]. The Néel
temperature has been found to increase due to reinforcement of the 3d–3d exchange
interactions, whereas the spin reorientation phenomenon occurs also at higher
temperature upon Co substitution for Fe [468].
Studying the transport properties of some R6 Fe11 Al3 compounds, about 4%
negative magnetoresistance has been reported in Nd6 Fe11 Al3 at 5 K in a field
of 6 T, but there is no magnetoresistance in La6 Fe11 Al3 [474]. An additional
magnetoresistance effect has been reported to occur at the metamagnetic transition
for R6 Fe13–x Al1+x with both R = Nd and La. Indeed, at the critical field, a 3–5%
resistivity drop has been found [475].
Hydrogen insertion can be performed in the crystal structure of the R6 M13 A
compounds. The maximum hydrogen content ranges from 13 to 20 depending on
the sample and authors [461, 471, 476–478]. Hydrogen has been reported to be
accommodated in four to six inequivalent interstitial sites of either tetrahedral or
octahedral type in Pr6 Fe13 AuD13 and Nd6 Fe13 GaD12 [477, 478], and it induces a
large and highly anisotropic deformation, being an order of magnitude greater along
c than along a [471]. The lattice parameters and main magnetic properties of the
R6 M13 AHx compounds are summarized in Table 25.
Contrary to the antiferromagnetic ground state of the R6 M13 A compounds, the
corresponding hydrides are all ferromagnetic for light R elements and have an
easy magnetization direction along the c-axis due to the Fe-sublattice anisotropy.
The effect of H insertion on the magnetization curves of Nd6 Fe13 A compounds
has been studied for the compounds with R = Au and Sn [471]. The magnetic
order of the hydrides has been estimated to about 450 K, just slightly above
that of the parent compounds [461]. Reference [476] gives a comparison of the
Mössbauer spectra for the R6 M13 A compounds and the corresponding hydrides for
A = Sb, Bi, and Cu and R = Pr and Nd. Studying Nd6 Fe13 Au and Nd6 Fe13 AuH13 ,
increases in hyperfine field and isomer shift are observed upon hydrogenation [473].
The ferromagnetic ground state of the hydrides has been also shown by neutron
diffraction investigation, and large magnetic moments have been reported for Fe
atoms in Pr6 Fe13 AuD13 compounds (Table 25). Complementary investigation of
the 119 Sn Mössbauer spectroscopy has been performed on both R6 M13 Sn and
Table 24 Magnetic moment in (μB ) as derived from neutron diffraction experiments at 1.5 K for some R6 Fe13–x Ax compounds
15 Metallic Magnetic Materials

Nd6 Fe13 Sn
Nd6 Fe13 Au Pr6 Fe13 Si Pr6 Fe12 Ga2 Nd6 Fe12 Ga2 1.5 K Pr6 Fe13 Sn Nd6 Fe13 Si Nd6 Fe13 Si Pr6 Fe13 Au Pr6 Fe13 AuD13
1.5 K 1.5 K 1.5 K 1.5 K [464]tilted 1.5 K 2K 150 K 1.5 K 2K
Site [470]planar [470]planar [479]planar [479]axial 10◦ /c [464]planar [463]planar [463]axial [462]basal [478]basal
R1 8f 3.4 2.8 2.6 2.97 3.06 2.91 3.1 2.4 2.6 2.7
R2 16l 2.5 2.7 2.4 2.67 2.45 2.73 2.5 1.7 2.6 1.7
Fe1 4d 2.6 1.8 2.1 2.4 2.0 2.3 2.8 3.0 0.8 2.2
Fe2 16k 2.6 2.0 2.1 2.4 2.67 2.57 2.6 2.9 2.6 1.7
Fe3 16l1 2.6 2.1 1.9 2.4 2.45 2.42 2.4 2.7 2.4 3.4
Fe4 16l2 2.6 2.1 1.9 2.4 2.53 2.54 1.8 2.0 2.2 1.6
755
756 J. P. Liu et al.

Table 25 Lattice parameters and magnetic properties of R6 M13–x Ax Hy (M=Fe, Co) compounds
Ms Ms
(μB /f.u.) (μB /f.u.)
R6 Fe13–x Ax Hy a (10−3 nm) c (10−3 nm) TC (K) 4K 293 K Ref.
La6 Fe11 Al3 H16 829.1 2604. 444 22.7 [461]
La6 Fe13 SnH14.5 819.9 2537.6 [480]
Pr6 Fe13 AgH17 827.0 2557.8 19.4 [471]
Pr6 Fe13 AuH17 820.4 2547.4 23.1 [471]
Pr6 Fe13 AuD13 817.4 2540.8 [477]
Pr6 Fe13 SiH14.7 812.3 2541.0 22.6 [471]
Pr6 Fe13 GeH13.9 812.8 2542.1 21.4 [471]
Pr6 Fe13 SnD12.4 818.9 2542.5 20.5 [471]
Pr6 Fe13 PbH13.1 819.9 2542.5 22.1 [471]
Pr6 Fe13 SbH13.1 816.3 2539.3 491 30.8 [476]
Pr6 Fe13 BiH13.8 818.5 2542.2 485 31.7 [476]
Nd6 Fe13 CuH20 817.7 2525.6 469 30.9 [476]
Nd6 Fe13 AgH18.0 819.8 2535 487 36.8 18.8 [471]
Nd6 Fe13 AuH16.6 808.4 2227 41.5 24.9 [461]
Nd6 Fe13 SiH14.7 810.3 2525.4 23.5 [461, 471]
Nd6 Fe13 SnH13.3 816.3 2524.5 22.8 [471]
Nd6 Fe13 GeH14.5 809.8 2526.6 22.8 [471]
Nd6 Fe13 GaH15 815.2 2521.0 458 40.1 [477]
Nd6 Fe13 PbH13.1 816.4 2526.7 22.6 [471]
Nd6 Fe13 SbH13.1 812.3 2522.0 31.9 [476]
Nd6 Fe13 BiH13.7 814.2 2526.6 33.1 [476]

Table 26 Lattice parameters and magnetic properties of some R6 Co13–x Ax compounds


Ms
R6 Co13–x Ax a (10−3 nm) c (10−3 nm) TN (K) TSR (K) (μB /f.u.)4 K Ref.
La6 Co13 In 810.2 2357.6 490 11.5 [457]
La6 Co13 Tl 808.7 2354.8 13.3 [457]
La6 Co13 Ge 806.2 2292.2 [466]
La6 Co13 Sn 809.6 2341.6 14.3 [457]
La6 Co13 Pb 810.1 2353.3 12.3 [457]
La6 Co13 Sb 809.7 2328.9 13.7 [457]
La6 Co13 Bi 811.4 2347.8 [455]

corresponding hydrides R = La, Pr, Nd, and M = Fe, Co [480]. This has confirmed
the exclusive location of the Sn atom on one unique crystal position.
Such δ phase crystal structure can also be obtained with Co instead of Fe. In
La6 Co13 A (A = In, Tl, Sn, Pb, Sb, Bi), the mean cobalt magnetic moment has been
estimated to about 1 μB /Co atom [457], the ordering temperature being about 490 K
for La6 Co13 In. Lattice parameters and some magnetic properties of R6 Co13–x Ax
compounds are given in Table 26.
15 Metallic Magnetic Materials 757

Sm-Co Permanent Magnets

The story of Sm-Co magnets can be traced back to the 1940s, when the study of
rare-earth metals was greatly accelerated because of advances in chemical sepa-
ration techniques developed in association with the Manhattan Project. Magnetic
properties of GdCo5 were first reported by Nesbit [481] in 1959 and Hubbard [482]
in 1960. However, the significance of their work was overlooked, perhaps due to the
high cost of Gd and low magnetization of GdCo5 .
In 1966, Hoffer and Strnat [483] reported the extremely large magnetocrystalline
anisotropy of YCo5 . They further predicted that YCo5 and most other RCo5
compounds were candidates for new high-performance magnet materials. The
significance of this discovery was immediately recognized, and extensive studies
followed to determine the magnetic properties of RCo5 (R = Y, Sm, Ce, La, Nd, and
mischmetal) [484]. The first YCo5 specimens made by Strnat’s group in 1966 had
(BH)max only around 8 kJ/m3 , and in 1967, the same group reported 41 kJ/m3 for
SmCo5 [485]. Further improvements in (BH)max for SmCo5 resulted in 147 kJ/m3
in 1968 by Velge and Buschow [322] and 159 kJ/m3 by Das [486], Benz, and Martin
[487]. Today, the best (BH)max of SmCo5 is around 199 kJ/m3 . Partial substitution
of Pr for Sm gave slightly enhanced magnetization and energy product.
R2 Co17 compounds possess higher magnetization and Curie temperatures than
RCo5 . However, it proved to be very difficult to develop useful coercivity in
Sm2 Co17 [488–490]. It took nearly 10 years for Sm2 (Co,Fe)17 -based magnets to
reach 240 kJ/m3 by using Cu and Zr substitution for Co, coupled with a long
and complex processing [491–493]. The best (BH)max of Sm2 (Co,Fe,Cu,Zr)17 -type
magnets is 271 kJ/m3 at room temperature.
Figure 15 shows the binary Sm-Co phase diagram [494–496]. There exist quite
a few intermetallic compounds in the Sm-Co system, where SmCo5 and Sm2 Co17
possess important technical significance. As shown in Fig. 16 [497], SmCo5 has
a hexagonal crystal structure (1:5H, space group: P6/mmm; prototype: CaCu5 ),
while Sm2 Co17 has a rhombohedral crystal structure (2:17R, space group: R 3 m;
prototype: Th2 Zn17 ) at room temperature and a hexagonal crystal structure (2:17H,
space group: P63 /mmc; prototype: Th2 Ni17 ) at high temperatures (1300–1340 ◦ C).
Though applications of Nd-Fe-B magnets have greatly expanded since the mid-
1980s, Sm-Co magnets are still important because of their excellent thermal stability
and better corrosion resistance. Sm-Co magnets can be used at temperatures up to
∼300 ◦ C owing to their high Curie temperatures.
The Sm-Co magnets are based on SmCo5 and Sm2 Co17 phases. The composition
of the SmCo5 magnet is slightly Sm-rich, as compared with its chemical stoichiom-
etry. Its microstructure is basically a featureless single phase. On the other hand, the
composition and microstructure of Sm2 Co17 -based magnets are more complicated.
Figure 17 shows the TEM microstructure of a Sm(Co0.715 Fe0.20 Cu0.06 Zr0.025 )7.1
magnet after sintering at 1200 ◦ C for 1 h, followed by a solid solution heat
treatment at 1180 ◦ C for 5 h, isothermal aging at 800 ◦ C for 40 h, and slow
cooling from 800 ◦ C to 400 ◦ C at 1 ◦ C per minute. Its intrinsic coercive force is
1200 kA/m.
758 J. P. Liu et al.

Weight Percent Samarium


0 10 20 30 40 50 60 70 80 90 100
1600
1495°C Co5+xSm

1400 aùb Co19Sm5


1325
1260°C L
1240°C
b Co17Sm2

1200 1200°C
bCo7Sm2
1074°C 1074°C
1100
Temperature °C

1000 (g Sm)
(aCo) 922°C
a Co7Sm2
a Co17Sm2

800 (bSm)
Co5-xSm

Co2Sm
734°C
Co3Sm

695°C
605°C 595°C
600 575°C
~64 ~83
422°C

400

Co4Sm9

CoSm3
(εCo) (aSm)
200

0
0 10 20 30 40 50 60 70 80 90 100
Co Atomic Percent Samarium Sm

Fig. 15 Binary Co-Sm equilibrium phase diagram

Fig. 16 Crystal structures of SmCo5 (H) (a), Sm2 Co17 (R) (b), and Sm2 Co17 (H) (c)

The typical microstructure of a sintered Sm2 (Co,Fe,Cu,Zr)17 -type magnet shows


grain micrometers in size, with a fine, nanometric cellular structure within each
grain. As shown in Fig. 17, the cellular structure consists of three coherent phases:
a 2:17 cell interior phase, a 1:5 cell boundary phase rich in Sm and Cu, and a
platelet phase rich in Zr. It is believed that the domain wall pinning in the 1:5 cell
15 Metallic Magnetic Materials 759

Fig. 17 TEM micrograph of a sintered Sm(Co,Fe,Cu,Zr)7.1 with i Hc = 1200 kA/m. 1, cell


interior; 2, cell boundary; 3, platelet phase

boundary phase is the origin of the high coercivity, while the high coercivity in
SmCo5 magnets is generally attributed to nucleation.
Sintered Sm-Co magnets are very hard and brittle; therefore, machining them
into the final shape and size is troublesome, especially for tiny magnetic parts.
This led to the development of bonded Sm-Co magnets [497], which are made
by consolidating a magnet powder with a polymer matrix. Thermosetting binders,
such as epoxy resin, are employed for use in compression-molded magnets, while
thermoplastic binders, like nylon, for injection-molded magnets, and elastomers
, such as rubber, are used for extruded magnets [498]. Table 27 lists magnetic
properties of some Sm-Co magnets.
A fundamental change takes place for the coercivity mechanism when the
grain size reduces from micrometer to nanometer range. Intrinsic coercivity
of 1240 kA/m was easily obtained after annealing a high-energy ball-milled
stoichiometric Sm2 Co17 specimen at 750 ◦ C for only 1 min [499]. Surfactant-
assisted high-energy ball milling has been used to produce anisotropic SmCo5
nanoflakes, and the subsequent magnetic alignment and compaction would yield
bulk, anisotropic, nanocrystalline SmCo5 magnets [500–503]. Alternatively, bulk,
anisotropic, nanograin SmCo5 magnets with coercivity of 795–3980 kA/m and
(BH)max of 88–135 kJ/m3 were synthesized by hot compacting the high-energy
ball-milled SmCo5 powder at 700 ◦ C, followed by hot deformation at 800–900 ◦ C
with a height reduction of 70–90% [501, 504].
760 J. P. Liu et al.

Table 27 Magnetic properties of some commercial Sm-Co magnets (TM stands for Co,Fe,Cu,Zr)
(BH)max
Magnets μ0 Mr (T) i Hc
(kA/m) B Hc(kA/m) (kJ/m3 ) Ref.
SmCo5 , (Sm,Pr)Co5 0.8–1.0 >1900 600–780 130–190 a,b
Sm2 TM17 0.9–1.2 >1900 710–840 180–255 a,b
Temp. compensated 0.5–0.75 >1900 400–600 64–120 a
SmCo5
Temp. compensated 0.8–0.95 >1900 480–720 80–175 a
Sm2 TM17
Bonded SmCo5 0.4–0.5 600–1600 240–520 32–65 [497],b
Bonded Sm2 TM17 0.6–0.8 400–1600 310–530 64–130 [497],b
a http://www.electronenergy.com/products/samarium-cobalt.htm
b http://www.western-magnet.com/product.php

In a nanograin rare-earth magnet material, the rare-earth content can be reduced


to below than its chemical stoichiometric composition, resulting in a hard/soft
nanocomposite magnet material. In nanocomposites, because of the hard/soft
interfacial exchange coupling, the direction of magnetization in the soft phase is
constrained by that in the hard phase and tends to be aligned in the same direction
as in the hard phase. The exchange interaction of magnetic moments at the hard/soft
interface is, in a way, like a spring, leading to the term “exchange spring.”
In recent years, many processing approaches were adopted trying to make
bulk, anisotropic, nanocomposite SmCo5 /α-Fe magnets, such as powder blending,
powder particle coating, and magnetic-field-assisted ball milling [502, 503, 505];
however, the magnetic performance of the bulk nanocomposite magnets is still to be
improved.

Nd-Fe-B Permanent Magnets

Less expensive iron-based rare-earth magnets having magnetic properties com-


parable to those of Sm-Co had long been desired, but no suitable compounds
were available until the early 1980s. Of the known Nd-Fe phases, Nd2 Fe17
features the highest Curie temperature, Tc , of only 300 K. Despite this apparent
obstacle, two lines of research led to the preparation of iron-based hard magnets
by 1983. One approach relied on rapid solidification, by melt spinning, of rare-
earth iron–metalloid alloys (e.g., Croat et al. [506, 507] and the other employed
traditional powder metallurgy (sintering) techniques (Sagawa et al. [508])). The
various alloy compositions differed substantially, but all contained a novel ternary
crystalline compound responsible for the magnetic properties. Initially identified as
R3 Fe16 B, R3 Fe20 B, R3 Fe20 B2 , R3 Fe21 B, or R5 Fe25 B3 , the correct stoichiometry,
R2 Fe14 B, and detailed crystal structure were soon established for the prototypical
representative Nd2 Fe14 B first by neutron powder diffraction analysis and confirmed
by two independent single-crystal X-ray investigations (see the review by Herbst
15 Metallic Magnetic Materials 761

[509]). From the scientific perspective, the identification of Nd2 Fe14 B spawned a
great deal of exploration for structural isomorphs, determination of their intrinsic
properties, and investigation of the mechanisms responsible for those properties. On
the technological side, magnets based on Nd2 Fe14 B continue to dominate the high-
performance magnet market. Larger energy products (BH)max in conjunction with
the cost advantage over Sm-Co have enabled progressive growth in the spectrum of
applications for Nd-Fe-B. Approximately 80,000 metric tons of Nd-Fe-B magnets
was produced in 2015 with forecasts of 120,000 tons in 2020.
Of the elemental, eight binary, and three ternary members of the Nd-Fe-B
phase diagram (Fig. 18), only Nd2 Fe14 B features sufficient magnetization and
magnetocrystalline anisotropy to qualify as a permanent magnet candidate. Its unit
cell, displayed in Fig. 19, is tetragonal (P42 /mnm space group, No. 136), contains
four formula units (68 ions), and comprises six crystallographically distinct iron
sites, two different rare-earth positions, and one boron site. There is an eight-layer
repeat structure perpendicular to the c-axis. All Nd and B ions and the four Fe(4c)
ions reside in the basal (z = 0) and z = ½ mirror planes. Between these planes, the
other Fe ions form three puckered nets. The Fe(16k1 ), Fe(16k2 ), Fe(8j1 ), and Fe(4e)
sites comprise two slightly distorted hexagonal meshes rotated by ∼30◦ with respect
to one another; they enclose a network of Fe(8j2 ) sites located above or below the
centers of hexagons in the neighboring layers.
Many similarities exist between Nd2 Fe14 B and other crystal structures. The
hexagonal Fe layers resemble the nets in the σ -phase occurring in the Fe-Cr and
Fe-Mo systems. Among features common to Nd2 Fe14 B and the hexagonal CaCu5
structure of SmCo5 is the centering of the rare-earth atoms within hexagonal prisms.
Further parallels can be identified since many rare-earth transition metal (RE-
TM) structures are derivatives of CaCu5 . For example, the rhombohedral Th2 Zn17
structure in which Sm2 Co17 and Nd2 Fe17 crystallize can be obtained from CaCu5
by appropriate ion substitutions.
A link to TM–metalloid systems is evident from the boron coordination in
Nd2 Fe14 B. As Fig. 20 shows, each boron occupies the center of a trigonal prism
formed by the three nearest Fe atoms above and the three below either the basal or
z = ½ planes. Connecting the Fe layers above and below the planes containing Nd
and B, the prisms clearly contribute to the stability of the structure. Such prisms
are fundamental to the structure of many TM–metalloid materials, crystalline (e.g.,
FeB, Fe3 C, Fe3 P) as well as amorphous.
Table 28 lists intrinsic properties of a few magnetic materials for comparison
purposes. The saturation magnetization Ms sets the theoretical upper limit (BH)theor
max
= (μ0 Ms )2 /4μ0 for the energy product, and the anisotropy field HA represents the
largest intrinsic coercivity Hci that can be achieved in a practical magnet. Commer-
cial magnets based on Nd2 Fe14 B, Pr2 Fe14 B, SmCo5 , Sm2 Co17 , Sm2 Fe17 N3 , and
ferrites are available. Although ferrites are characterized by (BH)max of at most
∼50 kJ/m3 , they are the least expensive hard magnets per unit energy product
and very broadly used. Pr2 Fe14 B and Nd2 Fe14 B exhibit similar properties, and
for that reason less abundant Pr is frequently not separated from rare-earth ores
and retained as a constituent of magnets primarily containing Nd2 Fe14 B. For low-
762 J. P. Liu et al.

Fig. 18 Nd-Fe-B room temperature phase diagram [510]

temperature applications, Pr2 Fe14 B-based magnets are of significance, however, due
to the fact that the spin reorientation in Nd2 Fe14 B below 135 K does not occur
in Pr2 Fe14 B. Elemental Fe is included in Table 28 since its large Ms implies an
max = 920 kJ/m . Unfortunately, that cannot be realized since HA
enormous (BH)theor 3

is much too small; it must be substantially greater than Ms . While Dy2 Fe14 B and
Tb2 Fe14 B are not of interest as permanent magnets by themselves owing to the
scarcity of Dy and even more so Tb, their large anisotropy fields make Dy and Tb
very useful as low-level substituents to increase the coercivity of Nd2 Fe14 B-based
magnets and to improve its temperature dependence.
Fabrication of a practical permanent magnet from a compound having suitable
intrinsic properties, principally those of Table 28, requires a specific processing
method with an associated starting composition. The energy product and coercivity
are crucially dependent on the resultant microstructure, which involves the geometry
and orientation of the crystallites of the compound and the character and distribution
of secondary phases that can control domain wall formation and motion and,
hence, the magnetization and demagnetization behavior. Several methods are in
commercial use for the preparation of Nd-Fe-B magnets.
15 Metallic Magnetic Materials 763

Fig. 19 Unit cell of


Nd2 Fe14 B, the prototypical
R2 Fe14 B crystal structure.
The c/a ratio is expanded to
accentuate the hexagonal Fe
nets (Herbst [511]).
(Reprinted with permission
from [511] by American
Physical Society)

Powder metallurgy processing involves a series of orient, press, and sinter


operations on finely divided alloy powders. Long employed to produce ferrite and
Sm-Co magnets, it is now also extensively used for Nd-Fe-B. Magnets prepared
by this means are simply designated as sintered. Details vary, but the process
can be summarized as follows (see, e.g., Buschow [513], for an in-depth review).
Starting alloy near the Nd15 Fe77 B8 composition, substantially richer in Nd and B
compared to stoichiometric Nd2 Fe14 B (Nd12 Fe82 B6 ) with, as in other methods,
possible replacement of some Nd by Dy to increase Hci , low-level Co substitution of
Fe to enhance Tc , other additions to improve corrosion resistance, etc., is induction
melted or strip cast and powdered to a particle size of ∼3 μm. The powder is
aligned in a magnetic field and pressed perpendicular to the alignment direction.
The compact is then sintered at ∼1370 K and usually further heat treated. Such
magnets are relatively complex multiphase systems and may contain, in addition
to Nd2 Fe14 B, Nd1.1 Fe4 B4 , Nd-rich phases, and Fe (as expected from the phase
764 J. P. Liu et al.

Fig. 20 Trigonal prism in the


Nd2 Fe14 B structure (Herbst
et al. [511]). (Reprinted with
permission from [511] by
American Physical Society)

Table 28 Intrinsic Compound μ0 Ms (T) μ0 HA (T) TC (K)


properties of various
Nd2 Fe14 B 1.60 7.3 585
magnetic materials: saturation
magnetization Ms and Pr2 Fe14 B 1.56 7.5 565
anisotropy field HA at 295 K Dy2 Fe14 B 0.71 15 598
and Curie temperature TC . Tb2 Fe14 B 0.70 22 620
HA is the external magnetic SmCo5 1.05 40 993
field required to rotate the Sm2 Co17 1.30 6.5 1100
magnetization from an easy
direction to a hard direction Sm2 Fe17 N3 1.54 21 749
[509, 512] Ferrite(e.g., BaFe12 O19 ) ∼0.5 ∼1.5 ∼700
Fe 2.15 0.05 1043

diagram, Fig. 18), Nd oxides and pores. As the micrograph of Fig. 21 indicates, the
average Nd2 Fe14 B grain size is ∼10 μm; Nd2 Fe14 B occupies ∼85% of the volume.
The Nd-rich phase occurring at grain boundary junctions and in thin intergranular
layers is vital, serving as a liquid phase sintering aid.
Another important processing method is the rapid solidification technique of melt
spinning, in which molten starting alloy is ejected through a crucible orifice onto
the rim of a substrate disc rotating with surface velocity vS . The cooling rate is
proportional to vS and can reach 106 K/s. Ribbon fragments typically 30–50 μm
thick, 1–3 mm wide, and 1 mm-1 cm long are produced. The microstructure and
15 Metallic Magnetic Materials 765

Fig. 21 X-ray composition


micrograph of a sintered
Nd15 Fe77 B8 magnet. T1 , T2 ,
and Nd denote Nd2 Fe14 B,
Nd11 Fe4 B4 , and a Nd-rich fcc
phase (Sagawa et al. [514])

magnetic properties are sensitive to the quench rate, with Hci and (BH)max each
featuring a maximum as a function of vS . The micrograph of Fig. 22 (a) shows
that optimally quenched ribbons consist of spheroidal Nd2 Fe14 B grains with an
average diameter of ∼30 nm, some two orders of magnitude smaller than that
of sintered magnets (cf. Fig. 21). Nd2 Fe14 B comprises ∼95% of the volume.
The brittle melt-spun flakes can be consolidated into bulk magnets by (i) cold
compaction with a binder; (ii) hot pressing, which yields full densification with
little change in the microstructure and only marginal alignment of the randomly
oriented grains in the as-cast ribbons; and (iii) die upsetting, a second hot press in a
larger diameter die that can lead to as much as ∼75% magnetic alignment and much
larger (BH)max . Remarkably, die upsetting transforms the spheroidal grains of the
hot-pressed precursor to platelets stacked transverse to the press direction with the
c-axes normal to the grain faces (Fig. 22b).
In the hydrogen disproportionation desorption recombination (HDDR) process,
starting Nd-Fe-B material is heated under flowing H2 gas and disproportionates
into Fe, Fe2 B, and Nd hydride. Continued heat treatment in vacuum desorbs the
hydrogen and, rather spectacularly, recombines the disproportionated products into
fine-grained, coercive Nd2 Fe14 B. HDDR is one means for preparing Nd-Fe-B
powder for the manufacture of anisotropic bonded magnets, and the method is also
promising for the recycling of sintered Nd-Fe-B material.
Nd-Fe-B magnets having energy products as large as ∼400 kJ/m3 (50 MGOe)
are available commercially, and research efforts to further enhance (BH)max toward
max [Nd2 Fe14 B] = 510 kJ/m (64 MGOe) continue. One
the ultimate limit (BH)theor 3

approach relies on reducing the amount of intergranular and secondary phases


so that sufficiently small Nd2 Fe14 B crystallites can be magnetically coupled and
aligned by the exchange interaction, improving (BH)max by increasing the rema-
nence (the magnetization at zero field). Nanocomposite structures offer a novel and
fascinating pathway for increasing (BH)max even further (see Poudyal and Liu [502]
for a review and Coey and Skomski [516] for early theoretical considerations). The
fundamental concept is to exchange couple grains of a high anisotropy phase such
as Nd2 Fe14 B with grains of a higher magnetization phase such as Fe, again leading
766 J. P. Liu et al.

Fig. 22 (a) Bright field transmission electron micrograph (TEM) of optimally quenched Nd-Fe-B
ribbon. The inset is a selected area diffraction pattern demonstrating random distribution of the
Nd2 Fe14 B grains. The arrow indicates the thin (∼2 nm) Nd-rich, B-deficient amorphous phase
of approximate composition Nd0.7 Fe0.3 (near the eutectic in the binary Nd-Fe phase diagram)
occupying the intergranular regions (Mishra [515]). (b) TEM of a die-upset Nd-Fe-B magnet
(courtesy R. K. Mishra). (With kind permission from Elsevier)

to greater remanence. Minimizing the amount of the former would enable progress
max [Fe] = 920 kJ/m (120 MGOe).
toward the ultimate practical limit (BH)theor 3

Magnetocaloric Intermetallics

Since the discovery of magnetocaloric effect (MCE) by Warburg in the late


nineteenth century, great efforts have been dedicated to theoretical and experimental
investigations of magnetic refrigeration materials [517]. A common feature of these
materials is the concurrent change of crystallographic and magnetic properties dur-
ing a phase transition. In other words, magnetic phase transitions always take place
along with a discontinuous change in lattice parameters and/or crystal symmetry.
For most of them, both magnetic field and pressure can drive the first-order phase
transitions. Thus, these materials also display barocaloric effect (BCE) induced
by hydrostatic pressure and/or elastocaloric effect (ECE) induced by uniaxial
stress, as predicted by some theoretical investigations. Recently, Mathur et al.
[518, 519] reviewed various caloric materials near ferroic phase transitions and
addressed the significance and importance of multicaloric effects (magnetocaloric,
barocaloric/elastocaloric, and electrocaloric effect) with respect to fundamental
issues and applied cooling techniques.
15 Metallic Magnetic Materials 767

In this section, we focus on the magnetocaloric intermetallics with a mag-


netostructural transformation, which also display considerable barocaloric/elas-
tocaloric effect due to the mutual dependencies between the lattice and spin order-
ing. These materials include FeRh, Gd5 (Si1−x Gex )4 , La(Fe,Si)13 , MnFe(P1−x Asx ),
NiMn-based Heusler alloys, hexagonal MM’X compounds, and Mn(As1−x Sbx ). All
the caloric effects can be characterized by entropy change
S and/or adiabatic
temperature change
Tad (Table 29).
FeRh crystallizes in the CsCl-type cubic structure and experiences an abrupt
volume expansion of
V/V ∼ 0.9% at the antiferromagnetic (AFM)–ferromagnetic
(FM) transition around room temperature, without the change in the structure
symmetry (space group). Both magnetic field and pressure can influence the AFM–
FM transition. Thus, large MCE, BCE, and ECE have been observed [520–522].
Gd5 (Six Ge1−x )4 compounds exhibit a giant MCE [523], which originates from a
coupled magnetostructural first-order transition with a volume change of
V/V ∼
0.4–1.0% depending on the compositions [524, 525]. Through tuning compositions
of Gd5 (Six Ge1−x )4 , giant MCE has been achieved in a very wide temperature range
from 20 K to 290 K [526]. Gd5 Si2 Ge2 also displays large barocaloric effect [527],
which is associated with the fact that the FM orthorhombic α-Gd5 Si2 Ge2 has lower
phase volume compared to paramagnetic (PM) monoclinic β-Gd5 Si2 Ge2 which
makes it possible to induce the magnetostructural transition by hydrostatic pressure.
The La(Fe1−x Six )13 alloys show giant MCE at a first-order transition around
200 K [528], which is ascribed to the magnetoelastic transition noting the negative
lattice expansion can be
V/V∼1.2–1.6%, though the crystal symmetry (space
group: Fm–3 c) remains unchanged [529, 530]. To shift the TC , Co or Mn doping
was introduced, and giant MCE was observed in a wide temperature range of 120–
330 K in La(Fe1−x Cox )11.9 Si1.1 and La(Fe1−x Mnx )11.7 Si1.3 [531, 532]. Moreover,
hydrogen [529, 533, 534] or carbon additions [535, 536] in various La(Fe13−x Six )
samples can effectively enhance TC , while |
S| remains large and nearly unchanged
for the former case and drops for the latter case. The giant magnetocaloric
compound LaFe11.33 Co0.47 Si1.2 also shows a large barocaloric effect [537].
MnFeP1−x Asx (0.15< × <0.66) alloys crystallize in hexagonal Fe2 P-type struc-
ture and undergo a first-order FM-PM transition around room temperature. The
symmetry (space group: P 62m) is unchanged, but the lattice shows an anomalous
expansion in the direction of the a and b axes and a contraction of the c-axis as
the FM phase is formed [538]. As a result, the volume expands by
V/V∼0.06%
with increasing temperature. For such a magnetoelastic transition, MnFeP1−x Asx
(0.15< × <0.66) displays large MCE in a wide temperature range from 200 K to
350 K on tuning the ratio of P/As [539]. Moreover, replacing As with nontoxic
Ge and Si can retain the first-order magnetoelastic transition and the giant magne-
tocaloric effect [540]. Applied pressure has some effects on the TC and MCE [541],
but no BCE or ECE was reported.
The FM Heusler alloy Ni2 MnGa with L21 structure undergoes a magnetostruc-
tural transformation from the FM austenitic to FM martensitic phase on cooling,
accompanied by a lattice contraction about
V/V∼1%. The different structural
symmetry leads to different magnetic properties; thus, a considerable inverse
Table 29 Summary of magnetocaloric (MCE), barocaloric (BCE), and elastocaloric effect (ECE) for the typical magnetocaloric intermetallics, FePh,
768

Gd5 (Si1−x Gex )4 , La(Fe,Si)13 , MnFe(P1−x Asx ), NiMn-based Heusler alloys, hexagonal MM’X compounds, and Mn(As1−x Sbx ), where
V/V denotes the
volume change of lattice across magnetostructural transformation,
S entropy change, and
Tad adiabatic temperature change (directly or indirectly obtained
by direct measurements or calculations)
Sample
V/V MCEa BCEa ECEa

S
Tad
S
Tad
S
Tad
0–
0–2T 5T 0–2T 0–5T
FeRh 0.9% 12 J/kgK I [522] 12 J/kgK I [522] −6 K I [520] – −12.5 J/kgK 8–10 K 13 J/kgKI [520] −5.17 K
316 K 316 K Direct 316 K C [522] C [520] 5.29 kbar I [520]

Fe49 Rh51 Fe49 Rh51 Fe49 Rh51 2.0kbar 2.5 kbar 316 K 5.29 kbar
316 K Direct Fe49 Rh51 Direct
Fe49 Rh51 316 K 316 K
Fe49 Rh51 Fe49 Rh51
Gd5 (Si1−x Gex )4 0.4∼ −14 J/kgK C [523] −18J/kgK C [523] 7.3 K C [523] 15.0 K C [523] −13 J/kgK −1.3 K – –
1.0% 276 K 276 K Indirect 276 K Indirect 276 K C [525] C [525]

Gd5 Si2 Ge2 Gd5 Si2 Ge2 Gd5 Si2 Ge2 Gd5 Si2 Ge2 2.9 kbar 276 K −2 kbar
Gd5 Si2 Ge2 Direct
276 K
Gd5 Si2 Ge2
La(Fe, Si)13 - −1.2∼ −14.4 J/kgK −19.4 J/kgK 8.1 K C [534] 12.1 K C [534] – – – –
based −1.6% C [528] 208 K C [528] 208 K Indirect 184 K Indirect 184 K
compounds LaFe11.4 Si1.6 LaFe11.4 Si1.6 LaFe11.7 Si1.3 LaFe11.7 Si1.3
−12 J/kgK C [530] −20 J/kgK C [530] – – 8.6 J/kgK I 2.3K I [537] – –
274 K 274 K [537] 2.1 kbar 2.1 kbar
LaFe11.2 Co0.7 Si1.1 LaFe11.2 Co0.7 Si1.1 250 K Direct 250 K
LaFe11.33 LaFe11.33
Co0.47 Si1.2 Co0.47 Si1.2
−16 J/kgK C [532] −20 J/kgK C [532]
155 K 155 K
LaFe11.5 Mn0.2 Si1.3 LaFe11.5 Mn0.2 Si1.3
J. P. Liu et al.
– −17 J/kgK C [533] 7.1 K C [534] 15.4 K C [534] – – – –
288 K Indirect Indirect
La(Fe11.5 Si1.5 )H1.3 184 K 184 K
LaFe11.7 Si1.3 H1.1 LaFe11.7 Si1.3 H1.1
– −12 J/kgK C [535] – – – – – –
250 K
La
(Fe11.6 Si1.4 )C0.6
MnFeP 0.06% −14 J/kgK C [539] −18 J/kgK C [539] – – – – – –
(As,Si,Ge) 304 K 304 K
15 Metallic Magnetic Materials

MnFeP0.45 As0.55 MnFeP0.45 As0.55


−16 J/kgK C [541] −41 J/kgK C [541] – – – – – –
292 K 292 K
MnFeP0.63 Si0.26 MnFeP0.63 Si0.26
Ge0.11 Ge0.11
NiMn-based −1% 7 J/kgK I [543] 19 J/kgKI [543] – – – –
Heusler alloys 307 K 307 K
NiMn37 Sn13 NiMn37 Sn13
– – −3.6∼−6.2K – −24.4 J/kgK – – 3.5KC
I [544] C [545] [546]
319 K 2.6 kbar 1 kbar
Direct 307 K Direct
Ni45.7 Mn36.3 Ni49.26 Mn36.08 300 K
In13 Co5.1 In14.66 Ni45.7
Mn36.6
In13.3 Co5
(Continued)
769
770

Table 29 (Continued)
Sample
V/V MCEa BCEa ECEa

S
Tad
S
Tad
S
Tad
0–2T 0–5T 0–2T 0–5T
Hexagonal −2.8∼ −11.6 J/kgK −30 J/kgK C [549] – – 52 J/kgK 18 K I [551] – –
MM’X −3.9% C [549] 322 K I [551] 3 kbar
compounds 322 K Mn0.96 Cr0.04 CoGe 3 kbar Indirect
Mn0.96 Cr0.04 CoGe 300 K 300 K
MnCoGe0.99 MnCoGe0.99
In0.01 In0.01
−7 J/kgK C [214] −19 J/kgK C [214] – – – – – –
266K 266K
MnNi0.77 Fe0.23 Ge MnNi0.77 Fe0.23 Ge
MnAs1−x Sbx −1.1∼ −32 J/kgK C [553] −32 J/kgK C [553] 5 K C [553] 13 K C [553] – – – –
−2.1% 318 K 318 K 318K 318 K
MnAs MnAs Indirect Indirect
MnAs MnAs
−26 J/kgK C [553] −26 J/kgK C [553]
309 K 309 K
MnAs0.95 Sb0.05 MnAs0.95 Sb0.05
−8.7 J/kgK −19 J/kgK C [555] – – – – – –
C [555] 315 K
315 K MnAs/- MnAs/GaAs(001)
GaAs(001)
−3.4 J/kgK −7 J/kgK C [555] – – – – – –
C [555] 315 K
315 K MnAs/- MnAs/GaAs(111)
GaAs(111)
a The superscripts I and C denote the inverse and conventional character of the caloric effect, respectively
J. P. Liu et al.
15 Metallic Magnetic Materials 771

MCE was initially reported in an off-stoichiometric Ni51.5 Mn22.7 Ga25.8 [542].


Furthermore, the Ni–Mn–Z Heusler alloys with excess Mn display metamagnetic
behavior. The large difference of saturated magnetization across the martensitic
transformation results in a large Zeeman energy μ0
M·H [203], which can drive
a structural transformation and cause the field-induced metamagnetic behavior,
thus leading to a large inverse
S, as observed in Ni50 Mn37 Sn13 [543] and
Ni–Mn–In–Co alloy [544]. Moreover, the metamagnetic alloys also display large
BCE [545] and ECE [546].
As members of the hexagonal MM’X family compounds, MnCoGe/MnNiGe-
based alloys undergo a giant negative thermal expansion across the magne-
tostructural transformation. The change of unit cell volume can be as large as

V/V∼2.68–3.9% [214, 547] depending on composition, indicating that a big


change of internal energy is involved. Many groups reported magnetocaloric effect
in MnCoGe/MnNiGe -based compounds with various doping including vacancies
[213, 214, 548–550]. Giant barocaloric effect was also observed [551]. The achieved
inverse
S under a moderate pressure of 3 kbar reaches 52 Jkg−1 K−1 , while the
calculated
Tad reaches 18 K, which exceeds that of most materials, including the
reported giant magnetocaloric effect driven by 5 T magnetic field that is available
only by superconducting magnets.
MnAs shows a large MCE [552] due to a magnetostructural transition from
PM hexagonal NiAs-type to an FM orthorhombic MnP-type structure with a large
anomaly decrease in volume of
V/V∼1.6%. The substitution of As with Sb in
MnAs1−x Sbx can tune the transition temperature and reduce the thermal hysteresis
[553]. Applied pressure has obvious effects on the TC and MCE [554]. Strain
engineering of the MCE was studied in MnAs epilayers [555]. The epilayer of MnAs
on different GaAs(111) and GaAs (100) substrates experiences different strain
states, which resulted in different
S. Such a strain engineering of MCE suggests
an interesting route of magnetocaloric applications in microelectronic circuitry as
well as spintronic devices.

Heavy-Fermion Compounds

Heavy-fermion compounds are intermetallics that upon cooling to liquid helium


temperature show an extremely strong renormalization of the electron mass. The
field of heavy-fermion physics started in 1975, when specific heat measurements
on CeAl3 uncovered an enormous specific heat linear in temperature, c = γ T,
with a Sommerfeld coefficient γ of 1620 mJ/molK2 that pointed to band-like 4f
electrons. Such a very large value of the Sommerfeld coefficient is the experimental
hallmark of a heavy-fermion compound [556, 557]. While for normal metals the γ -
value is of the order of 1–10 mJ/molK2 , for a heavy-fermion system the γ -value is
typically 100–1000 mJ/molK2 , two or three orders of magnitude larger. The large γ -
value finds a phenomenological description in Landau’s model of the Fermi liquid
(FL): due to electron–electron interactions, heavy electrons or quasiparticles are
formed at the Fermi surface with an effective electron mass m*/me = γ */γ , where
772 J. P. Liu et al.

γ * is the enhanced γ -value (me is the free electron mass). Thus, heavy-fermion
compounds allow investigation of the electron mass renormalization paradigm
processes. Around 1990, it was realized that in some heavy-fermion compounds,
Landau’s Fermi liquid breaks down, since a linear electronic term in the specific
heat is not attained, even at the lowest temperatures. Consequently, new ideas and
new theoretical models had to be developed to explain the non-Fermi liquid (NFL)
behavior [558, 559]. Nowadays, heavy-fermion compounds are considered to be
ideal laboratory tools for the investigation of quantum phase transitions, where non-
Fermi liquid behavior occurs right at the quantum critical point.
The archetypical heavy-fermion compounds are mostly intermetallic compounds
that contain the 4f -element cerium [560] or the 5f -element uranium [561]. Near
room temperature, the f -moments behave as local (weakly interacting) magnetic
moments. Upon cooling, below a characteristic temperature that is typically 10 K,
termed the coherence temperature, the heavy-fermion behavior sets in. The key
ingredient in the formation of the heavy-fermion state is the hybridization of the
f -states with the s, p, or d orbitals of the ligand non-f -electron atoms in the
compound. The hybridization delocalizes the f -electrons resulting in a narrow
electron band at the Fermi level. This in turn leads to the high electron density of
states that is probed by the electronic specific heat. The phenomenology is captured
by the relations

c Vm kB2 m ∗ kF 2NA π 2 kB2


γ∗ = lim with γ∗ = = N (EF ) .
T →0 T 32 3

Here, c is the molar specific heat, kF is the Fermi wavenumber, and N(EF ) is the
enhanced density of states per atom per spin direction at the Fermi level, EF (Vm is
the molar volume, kB is the Boltzmann constant, -h is the Planck constant divided by
2π, and NA is Avogadro’s number).
On the microscopic level, the understanding of the strongly correlated electron
state has evolved historically via the concepts of valence fluctuations, virtual bound
4f states, Kondo-lattice formation, and the competition between the Kondo effect
and the Ruderman–Kittel–Kasuya–Yosida (RKKY) interaction [562], to the fields
of magnetic quantum phase transitions (QPT), quantum critical points (QCP), and
non-Fermi liquid behavior [563].
Most heavy-fermion compounds exhibit the Kondo effect. In its most simple
form, the Kondo Hamiltonian describes the exchange interaction between a single
magnetic impurity with spin S and a conduction electron with spin s: HK = − 2J
s · S. For an exchange coupling parameter J < 0, the impurity spin is completely
screened at low temperature, and a Kondo singlet is formed. The binding energy
for this single-ion Kondo effect is kB TK ∝ e−1/J N (EF ) /N (EF ), where TK is the
Kondo temperature. Heavy-fermion materials represent a periodic lattice formed
of magnetic impurity atoms (e.g., Ce atoms), called a dense Kondo-lattice. Upon
cooling below room temperature, the single-ion Kondo effect is replaced, near the
coherence temperature, by complex electron correlations which may result in a
nonmagnetic correlated electron ground state. A second important energy scale in
15 Metallic Magnetic Materials 773

Fig. 23 Simplified schematic phase diagram for heavy-fermion compounds based on the competi-
tion between the Kondo effect and RKKY interaction [562]. The solid lines represent the variation
of TK and TRKKY as a function of the exchange parameter J. HF, heavy-fermion (gray); FL, Fermi
liquid (yellow). Magnetic ordering (blue) vanishes at the critical value of the exchange parameter
Jc . Heavy-fermion behavior is observed in the vicinity of Jc

Kondo-lattice compounds is provided by the interaction between the f -moments


mediated by the conduction electrons, the RKKY interaction. It has a predominantly
antiferromagnetic character and may give rise to an antiferromagnetic ground state.
The corresponding energy scale is kB TRKKY ∝ J2 N(EF ), where TRKKY is the RKKY
temperature. The competition between the on-site Kondo effect and the inter-site
RKKY interaction is described in the Doniach phase diagram [562] (Fig. 23). When
TRKKY > TK , an antiferromagnetic ground state prevails, while a nonmagnetic Fermi
liquid ground state surfaces when TK > TRKKY . The competition between both
phenomena gives rise to strong electron–electron interactions, and, consequently,
heavy-fermion materials are located in the Doniach diagram in the vicinity of the
critical exchange interaction Jc . The exchange parameter J may be considered as
a control parameter that can be tuned, for instance, by mechanical or chemical
pressure (doping), to drive the material through the magnetic–nonmagnetic phase
transition. In the limit T → 0, this is an example of a magnetic quantum phase
transition (QPT) in an itinerant electron system.
Quantum phase transitions are phase transitions that take place at zero temper-
ature. They are induced by a nonthermal control parameter δ, such as mechanical
or chemical pressure or magnetic field. Since at T = 0 K the thermal fluctuations
are negligible, the phase transition which takes place at a critical value of the
tuning parameter, δ c , is driven by quantum fluctuations associated with Heisenberg’s
uncertainty principle. The description of magnetic quantum phase transitions [563]
774 J. P. Liu et al.

Fig. 24 Generic phase diagram in the vicinity of a continuous magnetic quantum phase transition
(after Ref. [563]). (With kind permission from SciPris). Magnetic order is depressed as a function
of the control parameter δ and vanishes at the quantum critical point (QCP). Close to the magnetic
phase boundary (dark gray region), the critical behavior is classical. In the region between the
dashed lines, the system behaves quantum critically

in itinerant fermion systems is due to Hertz and Millis. It is characterized by a


diverging order parameter (correlation length ξ ) and a diverging relaxation time τ . In
contrast to a classical magnetic phase transition, in the quantum case, the dynamics
and statics are strongly coupled via the relation τ ∝ ξ z , where z is the dynamical
critical exponent. At finite temperature, different regimes of the behavior of ξ are
identified: the quantum disordered regime, where the Fermi liquid picture still
applies, and the quantum critical regime, where strong departures from the Fermi
liquid behavior occur. At the QCP, the Fermi liquid picture breaks down. The generic
phase diagram for magnetic quantum phase transitions is shown in Fig. 24.
Profound experimental studies revealed that the Hertz–Millis scenario does not
always give an adequate description of the QPT in heavy-electron compounds.
In recent years, other contrasting models have been developed, notably the local
moment scenario by Coleman and coworkers [564]. More recently, this picture has
been extended to a global phase diagram of magnetic heavy fermions [565].
Prominent heavy-fermion materials are listed in Table 30. Most of these are
uranium or cerium intermetallics, but selected ytterbium, praseodymium, and
transuranium compounds exhibit heavy-fermion behavior as well. A first kind of
heavy-fermion compounds are the so-called vegetable heavy-fermion system. These
materials do not show long-range order down to the lowest temperature investigated
(often in the milliKelvin range) but rather remain a Fermi liquid. A second group
consists of heavy-fermion antiferromagnets. Here the ordering temperature (Néel
temperature, TN ) is typically in the range 1–10 K. The ordered moments are reduced
with respect to the local moment value pointing to itinerant magnetic order.
Table 30 Basic properties of non-superconducting heavy-fermion compounds. The crystal structure, year of discovery, the γ *-value expressed in
J/molf-atom K2 , type of ground state, and Néel temperature, TN , are tabulated. Types are FL, Fermi liquid; AF, antiferromagnet; QCP, quantum critical point; and
NFL, non-Fermi liquid
Compound Structure Year γ *(J/mol·K2 ) Type TN (K) Remarks Ref.
CeAl3 Hexagonal 1975 1.6 FL – First heavy-fermion compound [569]
YbCuAl Hexagonal 1977 0.26 FL – [570]
CeCu6 Monoclinic 1984 1.6 FL – [571, 572]
CeRu2 Si2 Tetragonal 1985 0.35 FL – Metamagnetic transition at 8 T [573, 574]
CePtSi Tetragonal 1987 0.8 FL – [575]
15 Metallic Magnetic Materials

U2 Pt2 In Tetragonal 1994 4.2 FL – [576]


PrAg2 In Cubic 1996 6.4 FL – Heavy-fermion due to quadrupolar [577]
fluctuations
YbT2 Zn20 Cubic 2007 >0.5 FL – T = Fe, Co, Rh, Os, Ir [578]
U2 Zn17 Rhombohedral 1984 0.54 AF 9.7 [579]
UCd11 Cubic 1984 0.84 AF 5.0 [580]
NpBe13 Cubic 1984 0.9 AF 3.4 [581]
UCu5 Cubic 1985 0.25 AF 15 Additional ordering at 1.2 K [582]
CePb3 Cubic 1985 >0.2 AF 1.1 [583]
YbBiPt Cubic 1991 8.0 AF 0.4 Low carrier semimetal [584]
CePd2 Al3 Hexagonal 1992 0.38 AF 2.8 [585]
YbAgGe Hexagonal 2004 0.15 AF 0.65 [586]
CeCu5.9 Au0.1 Monoclinic 1994 – QCP – Local moment QPT [587]
Ce0.92 La0.08 Ru2 Si2 Tetragonal 1997 – QCP – Itinerant QPT [588]
YbRh2 Si2 Tetragonal 2000 – NFL 0.07 Field-tuned QCP [589]
YbRh2 (Si0.95 Ge0.05 )2 Tetragonal 2003 – QCP – Local moment QPT [590]
Ce3 Pd20 Si6 Cubic 2007 – NFL 0.31 Also quadrupolar ordering at TQP = [591]
0.5 K
775
776

Table 31 Basic properties of heavy-fermion superconductors. The crystal structure, year of discovery, γ *-value, type of ground state, superconducting
transition temperature, Tc , and the Néel temperature, TN , are tabulated. SC, superconductor at ambient pressure; SC(p), superconductor under pressure; AF,
antiferromagnet; FM, ferromagnet. For SC(p) compounds, the maximum Tc and corresponding pressure are listed. The γ *-value is expressed in J/molf-atom ·K2
Compound Structure Year γ * (J/mol·K2 ) Type Tc (K) TN (K) Remarks Ref.
CeCu2 Si2 Tetragonal 1979 1.0 SC 0.5 – First heavy-fermion superconductor; close to AF order [592, 593]
with TN = 0.8 K
UBe13 Cubic 1984 1.1 SC 0.85 – Two SC transitions when doped with Th [594, 595]
UPt3 Hexagonal 1984 0.45 SC 0.55 – Two SC transitions Tc + = 0.49 K and Tc − = 0.43 K; [596–598]
weak AF order at TN = 5 K
URu2 Si2 Tetragonal 1986 0.18 SC 0.8 – Hidden order at 17.5 K [599]
CeCoIn5 Tetragonal 2001 0.29 SC 2.3 – [600]
CeIrIn5 Tetragonal 2001 0.72 SC 0.4 – [601]
PrOs4 Sb12 Cubic 2002 0.35 SC 1.85 – Two SC transitions at 1.75 K and 1.85 K [602, 603]
Ce2 CoIn8 Tetragonal 2002 0.50 SC 0.4 – [604]
β-YbAlB4 Orthorhombic 2008 0.15 SC 0.08 – NFL compound [605]
Ce2 PdIn8 Tetragonal 2009 1.00 SC 0.65 – [606]
UNi2 Al3 Hexagonal 1991 0.12 AF, SC 1.0 4.2 Small moment AF [607]
UPd2 Al3 Hexagonal 1991 0.15 AF, SC 2.0 14.2 Large moment AF [608]
CePt3 Si Tetragonal 2004 0.39 AF, SC 0.75 2.2 Non-centrosymmetric [609]
structure
J. P. Liu et al.
Ce3 PtIn11 Tetragonal 2015 1.2 AF, SC 0.32 2 [610]
CeCu2 Ge2 Tetragonal 1992 0.2 AF, SC(p) 0.6510 GPa 4.1 [611, 612]
CePd2 Si2 Tetragonal 1996 0.06 AF, SC(p) 0.42.8 GPa 10 [613][614]
CeRh2 Si2 Tetragonal 1996 0.08 AF, SC(p) 0.350.8 GPa 36 [615]
CeIn3 Cubic 1997 0.14 AF, SC(p) 0.182.5 GPa 10.1 [616]
CeRhIn5 Tetragonal 2000 0.42 AF, SC(p) 2.11.9 GPa 3.8 [617]
CeRhSi3 Tetragonal 2005 0.12 AF, SC(p) 1.12.0 GPa 1.6 Non-centrosymmetric [618]
structure
CeIrSi3 Tetragonal 2006 0.11 AF, SC(p) 1.6 2.5 GPa 5 Non-centrosymmetric [619]
15 Metallic Magnetic Materials

structure
UGe2 Orthorhombic 1991 0.03 FM, SC(p) 0.81.2 GPa – Coexistence FM and SC; TCurie = 53 K [620]
URhGe Orthorhombic 2001 0.16 FM, SC 0.25 – Coexistence FM and SC; TCurie = 9.5 K [621]
UIr Monoclinic 2004 0.05 FM, SC(p) 0.12.7 GPa – Coexistence FM and SC; TCurie = 46 K [622]
UCoGe Orthorhombic 2007 0.06 FM, SC 0.6 – Coexistence FM and SC; TCurie = 3.0 K [623]
777
778 J. P. Liu et al.

Another fascinating aspect of heavy-fermion physics is the superconducting


ground state that develops out of a strongly interacting electron liquid. The discovery
in 1979 of superconductivity at Tc ∼ 0.5 K in the heavy-electron compound
CeCu2 Si2 (γ * ∼ 1000 mJ/molK2 ) was a surprise. Whereas magnetic interactions
that lead to the strongly renormalized Fermi liquid state are normally detrimental
for superconductivity, in CeCu2 Si2 , the 4f band-like electrons participate in the
superconducting state. The discovery of superconductivity in the heavy 5f -electron
compounds UBe13 and UPt3 initiated the whole new research field of unconven-
tional superconductivity [566]. In traditional superconductors, superconductivity
is explained by the standard Bardeen–Cooper–Schrieffer model in which spin-
singlet Cooper pairs are mediated by the exchange of phonons. However, for
heavy fermion: (i) the superconducting state is described by a gap function with
line or point nodes, rather than a full gap over the whole Fermi surface, and (ii)
Cooper pairs are mediated by (anti)ferromagnetic spin fluctuations rather than by
phonons. Nowadays, heavy-fermion materials provide a vast, fruitful playground
to investigate the interplay of magnetic order and complex superconducting order
parameters [567, 568].
An overview of the most important heavy-fermion superconductors is presented
in Table 31. These can be divided into three main groups: heavy-electron compounds
with (i) a superconducting ground state, (ii) coexistence of superconductivity (under
pressure) and antiferromagnetic order, and (iii) coexistence of superconductivity
(under pressure) and ferromagnetic order. UBe13 and UPt3 belong to the first group
and served as prototypes to shape the field of unconventional superconductivity.

References
1. Gutfleisch, O., Willard, M.A., Brück, E., Chen, C.H., Sankar, S., Liu, J.P.: Magnetic materials
and devices for the 21st century: stronger, lighter, and more energy efficient. Adv. Mater. 23,
821–842 (2011)
2. Ayers, J., Harris, V., Sprague, J., Elam, W., Jones, H.: On the formation of nanocrystals in the
soft magnetic alloy Fe 73.5 Nb 3 Cu 1 Si 13.5 B 9. Acta Mater. 46, 1861–1874 (1998)
3. Hono, K., Ping, D., Ohnuma, M., Onodera, H.: Cu clustering and Si partitioning in the early
crystallization stage of an Fe 73.5 Si 13.5 B 9 Nb 3 Cu 1 amorphous alloy. Acta Mater. 47,
997–1006 (1999)
4. McHenry, M.E., Willard, M.A., Laughlin, D.E.: Amorphous and nanocrystalline materials for
applications as soft magnets. Prog. Mater. Sci. 44, 291–433 (1999)
5. Inoue, A., Takeuchi, A.: Recent development and application products of bulk glassy alloys.
Acta Mater. 59, 2243–2267 (2011)
6. Willard, M.A., Daniil, M.: Nanocrystalline soft magnetic alloys two decades of progress. In:
Buschow, K.H.J. (ed.) Handbook of Magnetic Materials, vol. 21, pp. 173–342 (2013)
7. Herzer, G.: Chapter 3 Nanocrystalline soft magnetic alloys. In: Handbook of Magnetic
Materials, vol. 10, pp. 415–462. Elsevier (1997)
8. Willard, M.A., Daniil, M.: Nanostructured soft magnetic materials. In: Nanoscale Magnetic
Materials and Applications, pp. 373–397. Springer (2009)
9. Willard, M., Laughlin, D., McHenry, M., Thoma, D., Sickafus, K., Cross, J.O., et al.: Structure
and magnetic properties of (Fe0. 5Co0. 5) 88Zr7B4Cu1 nanocrystalline alloys. J. Appl. Phys.
84, 6773–6777 (1998)
15 Metallic Magnetic Materials 779

10. Yoshizawa, Y., Yamauchi, K., Yamane, T., Sugihara, H.: Common mode choke cores using
the new Fe-based alloys composed of ultrafine grain structure. J. Appl. Phys. 64, 6047–6049
(1988)
11. Suzuki, K., Kikuchi, M., Makino, A., Inoue, A., Masumoto, T.: Changes in microstructure
and soft magnetic properties of an Fe 86 Zr 7 B 6 Cu 1 amorphous alloy upon crystallization.
Mater. Trans. JIM. 32, 961–968 (1991)
12. Knipling, K., Daniil, M., Willard, M.: Nanocrystalline Fe88− 2xCoxNixZr7B4Cu1 alloys:
soft magnets for vehicle electrification technologies. J. Appl. Phys. 117, 172611 (2015)
13. Chen, C.: Magnetism and Metallurgy of Soft Magnetic Materials. Dover Publications, New
York (1986)
14. O’handley, R.C.: Modern Magnetic Materials: Principles and Applications, vol. 830622677.
Wiley, New York (2000)
15. M. Inc., Alloy Specification Sheets.
16. Matsumoto, H., Urata, A., Yamada, Y., Inoue, A.: Novel Fe97-x-y Px By Nb2 Cr1 glassy alloys
with high magnetization and low loss characteristics for inductor core materials. IEEE Trans.
Magn. 46, 373–376 (2010)
17. Mishima, T.: Nickel-aluminum steel for permanent magnets. Stahl und Eisen. 53, 79 (1931)
18. Cahn, J.W.: Magnetic aging of spinodal alloys. J. Appl. Phys. 34, 3581–3586 (1963)
19. Cahn, J.W.: Phase separation by spinodal decomposition in isotropic systems. J. Chem. Phys.
42, 93–99 (1965)
20. Iwama, Y., Takeuchi, M.: Spinodal decomposition in AlNiCo 8 magnet alloy. T. Jpn. I. Met.
15, 371–377 (1974)
21. Iwama, Y., Takeuchi, M., Iwata, M.: New determination of magnetic anisotropy constants of
Alnico magnet alloys. J. Phys. Colloq. 32, C1-556–C1-557 (1971)
22. Cronk, E.R.: Recent developments in high-energy Alnico alloys. J. Appl. Phys. 37, 1097–
1100 (1966)
23. McCaig, M.: Phase changes in high coercivity alloys. Z. Angew. Phys. 21, 66 (1966)
24. Stanek, M., Leonowicz, M.: Investigations of thermo-magnetic treatment of Alnico 8 alloy.
Arch. Metall. Mater. 55, 571–577 (2010)
25. McCurrie, R.: Ferromagnetic Materials: Structure and Properties Academic, vol. 4, (1994)
26. Zhou, L., Miller, M.K., Dillon, H., Palasyuk, A., Constantinides, S., McCallum, R.W., et al.:
Role of the applied magnetic field on the microstructural evolution in Alnico 8 alloys. Metall.
Mater. Trans. E. 1, 27–35 (2014)
27. Zhou, L., Miller, M.K., Lu, P., Ke, L., Skomski, R., Dillon, H., et al.: Architecture and
magnetism of Alnico. Acta Mater. 74, 224–233 (2014)
28. Liu, T., Li, W., Zhu, M., Guo, Z., Li, Y.: Effect of Co on the thermal stability and magnetic
properties of Alnico 8 alloys. J. Appl. Phys. 115, 17A751 (2014)
29. Sun, Y., Zhao, J., Liu, Z., Xia, W., Zhu, S., Lee, D., et al.: The phase and microstructure
analysis of Alnico magnets with high coercivity. J. Magn. Magn. Mater. 379, 58–62 (2015)
30. Hoffmann, A., StÄblein, H.: Investigations of high-coercivity Alnico alloys. IEEE Trans.
Magn. 6, 225–230 (1970)
31. Takeuchi, M., Iwama, Y.: Effects of titanium upon magnetic anisotropy and coercivity in
alnico magnet alloys. T. Jpn. I. Met. 17, 489–496 (1976)
32. Mason, J.J., Ashall, D.W., Dean, A.V.: The structure of Ni-Co-Al-Ti-Cu-Fe permanent
magnets. IEEE Trans. Magn. 6, 191–194 (1970)
33. Hao, S.M., Ishida, K., Nishizawa, T.: Role of alloying elements in phase decomposition in
Alnico magnet alloys. Metall. Trans. A. 16, 179–185 (1985)
34. Bronner, C., Haberer, J.-P., Planchard, E., Sauze, J.: Contribution to the study of the influence
of Nb on the magnetic properties of Alnico alloys containing 4.5–6.5 % Ti. Cobalt. 46, 15–22
(1970)
35. Pramanik, S., Rao, V., Mohanty, O.: Effect of niobium on the directional solidification and
properties of Alnico alloys. J. Mater. Sci. 28, 1237–1244 (1993)
36. Palmer, D., Shaw, S.: Production of columnar castings in high-coercivity magnet alloys
containing 5–10 Per Cent Titanium. Cobalt. 43, 63–72 (1969)
780 J. P. Liu et al.

37. Zhou, L., Tang, W., Ke, L., Guo, W., Poplawsky, J.D., Anderson, I.E., et al.: Microstructural
and magnetic property evolution with different heat-treatment conditions in an Alnico alloy.
Acta Mater. 133, 73–80 (2017)
38. Zlatkov, B., Bavdek, U., Nikolic, M., Aleksic, O., Danninger, H., Gierl, C., et al.: Magnetic
properties of Alnico 8 sintered magnets produced by powder injection moulding. PIM Int. 3,
58–63 (2009)
39. Anderson, I., Kassen, A., White, E., Zhou, L., Tang, W., Palasyuk, A., et al.: Novel pre-
alloyed powder processing of modified Alnico 8: Correlation of microstructure and magnetic
properties. J. Appl. Phys. 117, 17D138 (2015)
40. Song, C., Han, B., Li, Y.: A study on Alnico permanent magnet powders prepared by
atomization. Cailiao Kexue Yu Jishu(J. Mater. Sci. Technol.)(China). 20, 347–349 (2004)
41. Tang, W., Zhou, L., Kassen, A., Palasyuk, A., White, E., Dennis, K., et al.: New Alnico
magnets fabricated from pre-alloyed gas atomization powder through diverse consolidation
techniques. In: Proceedings of Conference INTERMAG, pp. 1–1 (2015)
42. Chu, W., Fei, W., Li, X., Yang, D., Wang, J.: Evolution of Fe–Co rich particles in Alnico 8
alloy thermomagnetically treated at 800◦ C. Mater. Sci. Tech. 16, 1023–1028 (2000)
43. Sun, X., Chen, C., Yang, L., Lv, L., Atroshenko, S., Shao, W., et al.: Experimental study on
modulated structure in Alnico alloys under high magnetic field and comparison with phase-
field simulation. J. Magn. Magn. Mater. 348, 27–32 (2013)
44. Iwama, Y., Takeuchi, M.: Spinodal decomposition in AlNiCo 8 magnet alloy. Trans. Jpn. Ins.
Met. 15, 371–377 (1974)
45. Slater, J.C.: Electronic Structure of Alloys. J. Appl. Phys. 8, 385–390 (1937)
46. Pauling, L.: The nature of the interatomic forces in metals. Phys. Rev. 54, 899 (1938)
47. Bozorth, R.: Atomic moments of ferromagnetic alloys. Phys. Rev. 79, 887 (1950)
48. Ikeda, K., Nakamichi, T., Yamada, T., Yamamoto, M.: Ferromagnetism in Fe2Sc with the
Hexagonal MgZn2-Type Structure. J. Phys. Soc. Jpn. 36, 611–611 (1974)
49. Nevitt, M., Kimball, C., Preston, R.: Variation of 57Fe isomer shift with atomic size in
Laves phases and associated quadrupole and magnetic hyperfine fields. In: Proceedings of
International Conference on Magnetism, Nottingham September 1964, p. 137 (1964)
50. Ikeda, K., Gschneider, K.: Disappearance of the heat capacity peak of Sc 3 In around the
Curie temperature in high magnetic fields. J. Magn. Magn. Mater. 22, 207–211 (1981)
51. Yamada, H., Tohyama, T., Shimizu, M.: Metamagnetic transition of ScCo2 and LuCo2.
J. Phys. F: Met. Phys. 17, L163 (1987)
52. Acker, F., Fisk, Z., Smith, J.L., Huang, C.Y.: Enhanced paramagnetism of TiBe2 and
ferromagnetic transitions in TiBe2-xCux. J. Magn. Magn. Mater. 22, 250–256 (1981)
53. Brückner, W., Kleinstück, K., Schulze, G.E.R.: Mössbauer study of the laves phase Ti1-
xFe2+x. Phys. Status Solidi A. 1, K1–K4 (1970)
54. Brückner, W., Perthel, R., Kleinstück, K., Schulze, G.E.R.: Magnetic Properties of ZrFe2 and
TiFe2 within Their Homogeneity Range. Phys. Status Solidi B. 29, 211–216 (1968)
55. Aoki, Y.: Magnetic properties of the Intermetallic Compound with the Cu3Au-type structure
in cobalt-titanium alloy system. J. Phys. Soc. Jpn. 28, 1451–1456 (1970)
56. Aoki, Y., Nakamichi, T., Yamamoto, M.: Magnetic behavior of two laves phases in cobalt-
titanium alloy system. J. Phys. Soc. Jpn. 21, 565–566 (1966)
57. Bozorth, R.: Ferromagnetism (The Bell Telephone Laboratories Series). Van Nostrand, New
York (1951)
58. Read, D.A., Thomas, E.H., Forsythe, J.B.: Evidence of itinerant electron ferromagnetism in
sigma phase alloys. J. Phys. Chem. Solids. 29, 1569–1572 (1968)
59. Nevitt, M.V., Aldred, A.T.: Ferromagnetism in V-Fe and Cr-Fe alloys. J. Appl. Phys. 34, 463–
468 (1963)
60. Nevitt, M.V.: Curie Temperatures of Binary and Ternary Sigma Phases. University of Illinois
at Urbana-Champaign (1954)
61. Wolcott, N., Falge Jr., R.: Ferromagnetism of CrBe $ sub 1$. National Bureau of Standards,
Washington, DC (1968)
15 Metallic Magnetic Materials 781

62. Wachtel, E., Bartelt, C.: Measurements of the susceptibility in the chromium-manganese-
system. Z. Metallkd. 55, 29–36 (1964)
63. Martin, J., Downie, D.: Heat capacities of transition-metal alloys 1. Sigma-phase alloy Co
0.435 Cr 0.565. J. Chem. Thermodyn. 15, 691–699 (1983)
64. Kussmann, A., Muller, K., Raub, E.: Thermomagnetic measurements on alloys of the
platinum-group metals with chromium. Zeitschrift für Metallkunde. 59, 859–863 (1968)
65. Pickart, S., Nathans, R.: Neutron diffraction investigation of Pt-based alloys of the first
transition series. J. Appl. Phys. 34, 1203–1204 (1963)
66. Bauer, A., Regnat, A., Blum, C.G., Gottlieb-Schönmeyer, S., Pedersen, B., Meven, M., et al.:
Low-temperature properties of single-crystal CrB 2. Phys. Rev. B. 90, 064414 (2014)
67. Schoop, L., Hirschberger, M., Tao, J., Felser, C., Ong, N., Cava, R.: Paramagnetic to
ferromagnetic phase transition in lightly Fe-doped Cr 2 B. Phys. Rev. B. 89, 224417 (2014)
68. Coldea, M., Crişan, M., Néda, A., Pop, I.: NMR, magnetic susceptibility and specific heat of
CrAl 7. J. Phys. Chem. Solids. 35, 1095–1098 (1974)
69. Susner, M.A., Parker, D.S., Sefat, A.: Importance of doping and frustration in itinerant Fe-
doped Cr 2 Al. J. Magn. Magn. Mater. 392, 68–73 (2015)
70. Lee, C.-G., Youn, K.-T., Fukamichi, K.: Antiferromagnetic Cr-Ga layers with a high Neel
temperature for biasing spin valves. IEEE Trans. Magn. 36, 2902–2904 (2000)
71. Ohsugi, I.J., Kojima, T., Nishida, I.A.: Temperature dependence of the magnetic susceptibility
of a CrSi 2 single crystal. Phys. Rev. B. 42, 10761 (1990)
72. Hsu, H.-F., Tsai, P.-C., Lu, K.-C.: Single-crystalline chromium silicide nanowires and their
physical properties. Nanoscale Res. Lett. 10, 1–8 (2015)
73. Ghimire, N., McGuire, M.A., Parker, D.S., Sales, B.C., Yan, J.-Q., Keppens, V., et al.:
Complex itinerant ferromagnetism in noncentrosymmetric Cr 11 Ge 19. Phys. Rev. B. 85,
224405 (2012)
74. Browne, J., Liddell, P., Street, R., Mills, T.: An investigation of the antiferromagnetic
transition of CrN. Phys. Status Solidi A. 1, 715–723 (1970)
75. Wu, W., Cheng, J., Matsubayashi, K., Kong, P., Lin, F., Jin, C., et al.: Superconductivity in
the vicinity of antiferromagnetic order in CrAs. Nat. Commun. 5 (2014)
76. Shinohara, T., Watanabe, H.: Nuclear Magnetic Resonance in Cr 3 As 2. J. Phys. Soc. Jpn.
21, 2076–2076 (1966)
77. Ishimoto, K., Okonogi, M., Ohoyama, K., Nakajima, K., Ohashi, M., Yamauchi, H., et al.:
Anisotropic exchange interaction in Cr 2 As. Phys. B. 213, 336–338 (1995)
78. Snow, A.: Magnetic moment orientation and thermal expansion of antiferromagnetic CrSb.
Rev. Mod. Phys. 25, 127 (1953)
79. Shimada, K., Saitoh, T., Namatame, H., Fujimori, A., Ishida, S., Asano, S., et al.: Photoemis-
sion study of itinerant ferromagnet Cr 1− δ Te. Phys. Rev. B. 53, 7673 (1996)
80. Matsui, M., Ido, T., Sato, K., Adachi, K.: Ferromagnetism and Antiferromagnetism in Co–Mn
Alloy. J. Phys. Soc. Jpn. 28, 791–791 (1970)
81. Acet, M., John, C., Wassermann, E.: Magnetism and structural stability in CoMn alloys.
J. Appl. Phys. 70, 6556–6558 (1991)
82. Paoletti, A., Ricci, F., Passari, L.: Magnetization and State of Order in MnNi3. J. Appl. Phys.
37, 3236–3239 (1966)
83. Coey, J.M.: Magnetism and Magnetic Materials. Cambridge University Press, Cambridge
(2010)
84. Hori, T., Nakagawa, Y., Sakurai, J.: Magnetization and magnetic structure of Mn-Zn and Mn-
Zn-Ga alloys of CsCl-type structure. J. Phys. Soc. Jpn. 24, 971–976 (1968)
85. Leavey, C., Stewart, J., Rainford, B., Hillier, A.: Magnetic ground states and spin dynamics
of β-Mn1− xRux alloys. J. Phys.: Condens. Matter.19, 145288 (2007)
86. Krén, E., Kádár, G., Tarnóczi, T.: Atomic and magnetic order in Mn 2 Pd 3. Phys. Lett. A. 25,
56–57 (1967)
87. Caron, L., Hudl, M., Höglin, V., Dung, N., Gomez, C.P., Sahlberg, M., et al.: Magnetocrys-
talline anisotropy and the magnetocaloric effect in Fe 2 P. Phys. Rev. B. 88, 094440 (2013)
782 J. P. Liu et al.

88. Cadeville, M.: Proprietes magnetiques des diborures de manganese et de chrome: MnB 2 et
CrB 2. J. Phys. Chem. Solids. 27, 667–670 (1966)
89. Neov, S., Legrand, E.: Neutron diffraction study of the magnetic structure of Mn3B4. Phys.
Status Solidi B. 49, 589–596 (1972)
90. Lundquist, N., Myers, H., Westin, R.: The paramagnetic properties of the monoborides of V,
Cr, Mn, Fe, Co and Ni. Philos. Mag. 7, 1187–1195 (1962)
91. Park, J., Hong, Y., Bae, S., Lee, J., Jalli, J., Abo, G., et al.: Saturation magnetization and
crystalline anisotropy calculations for MnAl permanent magnet. J. Appl. Phys. 107, 09A731
(2010)
92. Bither, T.A., Cloud, W.H.: Magnetic tetragonal δ phase in the Mn–Ga binary. J. Appl. Phys.
36, 1501–1502 (1965)
93. Tsuboya, I., Sugihara, M.: Magnetic properties of ζ phase in Mn-Ga system. J. Phys. Soc. Jpn.
18, 1096–1096 (1963)
94. Tsuboya, I., Sugihara, M.: The magnetic properties of ε phase in Mn–Ga system. J. Phys.
Soc. Jpn. 18, 143–143 (1963)
95. Kharel, P., Huh, Y., Al-Aqtash, N., Shah, V., Sabirianov, R., Skomski, R., et al.: Structural
and magnetic transitions in cubic Mn3Ga. J. Phys.: Condens. Matter. 26, 126001 (2014)
96. Zhang, Q., Li, D., Cui, W., Li, J., Zhang, Z.: Magnetic properties and spin-glass-like behavior
in stoichiometric Mn (3) In compound. J. Appl. Phys. 106, 113915 (2009)
97. Karen, P., Fjellvag, H., Kjekshus, A., Andresen, A.: On the phase relations and structural and
magnetic properties of the stable manganese carbides Mn23C6, Mn5C2 and Mn7C3. Acta
Chem. Scand. 45, 549–557 (1991)
98. Mühlbauer, S., Binz, B., Jonietz, F., Pfleiderer, C., Rosch, A., Neubauer, A., et al.: Skyrmion
lattice in a chiral magnet. Science. 323, 915–919 (2009)
99. Yamada, N., Maeda, K., Usami, Y., Ohoyama, T.: Magnetic properties of intermetallic
compound Mn11Ge8. J. Phys. Soc. Jpn. 55, 3721–3724 (1986)
100. Kim, Y., Kim, E.J., Choi, K., Han, W.B., Kim, H.-S., Yoon, C.S.: Magnetocaloric effect of
Mn 5+ xGe 3− x alloys. J. Alloys Compd. 620, 164–167 (2015)
101. Timoschuk, V., Rozenberg, E.: Magnetic phase transitions of ζ-Mn 5 Ge 2 in magnetic fields
up to 25 T. Solid State Commun. 78, 531–534 (1991)
102. Wachtel, E., Henig, E.-T.: Magnetic properties of germanium and germanium- manganese
alloys in the solid and liquid states. Z. Metallkd. 60, 316–321 (1969)
103. Yasukōchi, K., Kanematsu, K., Ohoyama, T.: Magnetic Properties of Intermetallic Com-
pounds in Iron-Germanium System: Fe1. 67Ge and FeGe2. J. Phys. Soc. Jpn. 16, 429–433
(1961)
104. Duan, T., Ren, W., Liu, W., Li, S., Liu, W., Zhang, Z.: Magnetic anisotropy of single-
crystalline Mn3Sn in triangular and helix-phase states. Appl. Phys. Lett. 107, 082403 (2015)
105. Mazet, T., Recour, Q., Malaman, B.: Neutron diffraction and S 119 n Mössbauer spectroscopy
study of Mn 3 Sn 2. Phys. Rev. B. 81, 174427 (2010)
106. Felcher, G.: Magnetic structure of MnP. J. Appl. Phys. 37, 1056–1058 (1966)
107. Bacmann, M., Fruchart, D., Chenevier, B., Fruchart, R., Puertolas, J., Rillo, C.: Magnetic
phase diagram of the (Fe 1− x Mn x) 2 P system. J. Magn. Magn. Mater. 83, 313–314 (1990)
108. Leitão, J., Xinmin, Y., Caron, L., Brück, E.: Magnetostructural study of the (Mn, Fe) 3 (P, Si)
system. J. Alloys Compd. 520, 52–58 (2012)
109. Nascimento, F.C., dos Santos, A.O., de Campos, A., Gama, S., Cardoso, L.P.: Structural and
magnetic study of the MnAs magnetocaloric compound. Mater. Res. 9, 111–114 (2006)
110. Kanomata, T., Goto, T., Ido, H.: Magnetic phase transitions in the Fe2As-Mn2As system.
J. Phys. Soc. Jpn. 43, 1178–1184 (1977)
111. Takei, W., Cox, D., Shirane, G.: Magnetic structures in the MnSb-CrSb system. Phys. Rev.
129, 2008 (1963)
112. Shirakawa, K., Ido, H.: Magnetic transition of the Mn2Sb-Mn2As system. J. Phys. Soc. Jpn.
40, 666–673 (1976)
113. Adachi, K., Sato, K., Takeda, M.: Magnetic properties of some compounds with pyrite
structure. J. Appl. Phys. 39, 900–900 (1968)
15 Metallic Magnetic Materials 783

114. Massalski, T., Okamoto, H., Subramanian, P., Kacprzak, L.: Binary Alloys Phase Diagrams,
vol. II. ASM International, Materials Park (1990) 2001
115. Brando, M., Duncan, W., Moroni-Klementowicz, D., Albrecht, C., Grüner, D., Ballou,
R., et al.: Logarithmic Fermi-liquid breakdown in NbFe 2. Phys. Rev. Lett. 101, 026401
(2008)
116. Nikitin, S., Myalikgulyev, G., Tishin, A., Annaorazov, M., Asatryan, K., Tyurin, A.: The
magnetocaloric effect in Fe 49 Rh 51 compound. Phys. Lett. A. 148, 363–366 (1990)
117. Klemmer, T., Hoydick, D., Okumura, H., Zhang, B., Soffa, W.: Magnetic hardening and
coercivity mechanisms in L1 0 ordered FePd ferromagnets. Scr. Metall. Mater. 33, 1793–
1805 (1995)
118. Ikeda, K.: Ferromagnetism in hexagonal and cubic Fe 2 Hf compounds (1977)
119. Duijn, H., Brück, E., Menovsky, A., Buschow, K., De Boer, F., Coehoorn, R., et al.: Magnetic
and transport properties of the itinerant electron system Hf1− xTaxFe2. J. Appl. Phys. 81,
4218–4220 (1997)
120. Liu, Y., Jiang, Y., Zhang, X., Wang, Y., Zhang, Y., Liu, H., et al.: Structural and magnetic
properties of the ordered FePt 3, FePt and Fe 3 Pt nanoparticles. J. Solid State Chem. 209,
69–73 (2014)
121. Zeng, H., Li, J., Liu, J.P., Wang, Z.L., Sun, S.: Exchange-coupled nanocomposite magnets by
nanoparticle self-assembly. Nature. 420, 395–398 (2002)
122. Kakeshita, T., Takeuchi, T., Fukuda, T., Tsujiguchi, M., Saburi, T., Oshima, R., et al.: Giant
magnetostriction in an ordered Fe3Pt single crystal exhibiting a martensitic transformation.
Appl. Phys. Lett. 77, 1502–1504 (2000)
123. Kneller, E., Khan, Y.: The phase Fe sub 2 B. Z. Metallkd. 78, 825–835 (1987)
124. Dobrzynski, L., Petrillo, C., Sacchetti, F.: Iron magnetic moments and spin-density asphericity
in Fe 3 (Al x Si 1− x) alloys. Phys. Rev. B. 42, 1142 (1990)
125. Kawamiya, N., Adachi, K.: Magnetic structure of Fe 3 Ga studied by neutron diffraction.
Trans. Jpn. Inst. Met. 23, 296–302 (1982)
126. Taheri, P., Barua, R., Hsu, J., Zamanpour, M., Chen, Y., Harris, V.: Structure, magnetism,
and magnetostrictive properties of mechanically alloyed Fe 81 Ga 19. J. Alloys Compd. 661,
306–311 (2016)
127. Mendez, J., Ekuma, C., Wu, Y., Fulfer, B., Prestigiacomo, J., Shelton, W., et al.: Competing
magnetic states, disorder, and the magnetic character of Fe 3 Ga 4. Phys. Rev. B. 91, 144409
(2015)
128. Kobeissi, M., Hutchings, J., Appleyard, P., Thomas, M., Booth, J.: Mössbauer studies of the
alloys Fe3Ga4,(Fe1-xTix) 3Ga4 and (Fe1-yCry) 3Ga4. J. Phys.: Condens. Matter. 11, 6251
(1999)
129. Matsushita, M., Matsushima, Y., Ono, F.: Anomalous structural transformation and mag-
netism of Fe–Ga alloys. Phys. B. 405, 1154–1157 (2010)
130. Sluchanko, N., Glushkov, V., Demishev, S., Menovsky, A., Weckhuysen, L., Moshchalkov,
V.: Crossover in magnetic properties of FeSi. Phys. Rev. B. 65, 064404 (2002)
131. Adams, T., Mühlbauer, S., Neubauer, A., Münzer, W., Jonietz, F., Georgii, R., et al.: Skyrmion
lattice domains in Fe1− xCoxSi. J. Phys. Conf. Ser., 032001 (2010)
132. Yasukōchi, K., Kanematsu, K., Ohoyama, T.: Magnetic properties of intermetallic compounds
in manganese-Tin system: Mn3. 67Sn, Mn1. 77Sn, and MnSn2. J. Phys. Soc. Jpn. 16, 1123–
1130 (1961)
133. Yu, X., Kanazawa, N., Onose, Y., Kimoto, K., Zhang, W., Ishiwata, S., et al.: Near room-
temperature formation of a skyrmion crystal in thin-films of the helimagnet FeGe. Nat. Mater.
10, 106–109 (2011)
134. Felcher, G., Jorgensen, J.: Magnetic structures of monoclinic FeGe. J. Phys. C: Solid State
Phys. 16, 6281 (1983)
135. Beckman, O., Carrander, K., Lundgren, L., Richardson, M.: Susceptibility measurements and
magnetic ordering of hexagonal FeGe. Phys. Scr. 6, 151 (1972)
136. Drijver, J., Sinnema, S., Van der Woude, F.: Magnetic properties of hexagonal and cubic
Fe3Ge. J. Phys. F: Met. Phys. 6, 2165 (1976)
784 J. P. Liu et al.

137. Venturini, G., Malaman, B., Le Caër, G., Fruchart, D.: Low-temperature magnetic structure
of FeSn 2. Phys. Rev. B. 35, 7038 (1987)
138. Häggström, L., Ericsson, T., Wäppling, R., Chandra, K.: Studies of the magnetic structure of
FeSn using the Mössbauer effect. Phys. Scr. 11, 47 (1975)
139. Malaman, B., Fruchart, D., Le Caër, G.: Magnetic properties of Fe3Sn2. II. Neutron
diffraction study (and Mossbauer effect). J. Phys. F.: Metal Phys. 8, 2389 (1978)
140. Sales, B.C., Saparov, B., McGuire, M.A., Singh, D.J., Parker, D.S.: Ferromagnetism of Fe3Sn
and alloys. Sci. Rep. 4 (2014)
141. Felcher, G., Smith, F., Bellavance, D., Wold, A.: Magnetic structure of iron monophosphide.
Phys. Rev. B. 3, 3046 (1971)
142. Andersson, Y., Rundqvist, S., Beckman, O., Lundgren, L., Nordblad, P.: Properties of Fe2 P
crystals prepared from a liquid copper medium. Phys. Status Solidi A. 49, K153–K156 (1978)
143. Broddefalk, A., Granberg, P., Nordblad, P., Liu, H.-p., Andersson, Y.: Magnetocrystalline
anisotropy of (Fe1-xCox) 3P. J. Appl. Phys. 83, 6980–6982 (1998)
144. Selte, K., Kjekshus, A., Andresen, A.: Magnetic structure and properties of FeAs. Acta Chem.
Scand. 26, 3101–3113 (1972)
145. Önnerud, P., Andersson, Y., Tellgren, R., Ericsson, T., Nordblad, P., Krishnamurthy, A., et al.:
A magnetic structure investigation of tetragonal (Fe 1− x Co x) 2 As. J. Magn. Magn. Mater.
147, 346–354 (1995)
146. Petrovic, C., Kim, J.W., Bud’ko, S.L., Goldman, A., Canfield, P.C., Choe, W., et al.:
Anisotropy and large magnetoresistance in the narrow-gap semiconductor FeSb 2. Phys. Rev.
B. 67, 155205 (2003)
147. Kom˛edera, K., Jasek, A., Błachowski, A., Ruebenbauer, K., Krztoń-Maziopa, A.: Magnetic
anisotropy in FeSb studied by 57 Fe Mössbauer spectroscopy. J. Magn. Magn. Mater. 399,
221–227 (2016)
148. Buschow, K.: Intermetallic compounds of rare-earth and 3d transition metals. Rep. Prog.
Phys. 40, 1179 (1977)
149. Handbook of Magnetic Materials, vol. 4, p. 211, Elseviers Science (1988)
150. Fujii, H., Pourarian, F., Wallace, W.: Appearance of spontaneous ferromagnetism in non-
stoichiometric ZrCo 2. J. Magn. Magn. Mater. 24, 93–96 (1981)
151. Pareti, L., Solzi, M., Paoluzi, A.: Magnetocrystalline anisotropy of the 3d sublattice in the
cubic intermetallic system Zr6Co23− xMx (M= Fe, Ni). J. Appl. Phys. 73, 2941–2947 (1993)
152. Burzo, E., Grössinger, R., Hundegger, P., Kirchmayr, H., Krewenka, R., Mayerhofer, O., et al.:
Magnetic properties of ZrCo5. 1− xFex alloys. J. Appl. Phys. 70, 6550–6552 (1991)
153. Buschow, K.: Differences in magnetic properties between amorphous and crystalline alloys.
J. Appl. Phys. 53, 7713–7716 (1982)
154. Crangle, J., Parsons, D.: The magnetization of ferromagnetic binary alloys of cobalt or nickel
with elements of the palladium and platinum groups. Proc. Roy. Soc., 509–519 (1960)
155. Shen, Y., Turgut, Z., Horwath, J., Huang, M.: Bulk nanocomposite LaCo5/LaCo13 magnets.
J. Appl. Phys. 109, 07A765 (2011)
156. Franse, J., Gerdsdorf, R.: Magnetic properties of metals-3d, 4d and 5d elements, alloys and
compounds (Landolt-Börnstein, New Series vol 19a) ed HPJ Wijn. Springer, Berlin (1986)
157. Menzinger, F., Paoletti, A.: Magnetic moments and unpaired-electron densities in Co Pt 3.
Phys. Rev. 143, 365 (1966)
158. Cadeville, M., Dahmani, C., Kern, F.: Magnetism and spatial order in Ni-Pt and Co-Pt alloys.
J. Magn. Magn. Mater. 54, 1055–1056 (1986)
159. Sanchez, J., Mora-Loṕez, J., Leroux, C., Cadeville, M.: Chemical and magnetic ordering in
CoPt. J. Phys. Colloques. 49, C8-107–C8-108 (1988)
160. Onnerud, P., Andersson, Y., Tellgren, R., Nordblad, P.: The ferromagnetic structure of
hexagonal (Fe1-xCox) 2As. Solid State Commun. 101, 271–275 (1997)
161. Krumbügel-Nylund, A.: Thesis, Orsay, (1974)
162. Paccard, D., Pauthenet, R.: Proprietes Cristallographiques Et Magnetiques Des Alliages De
Formule TNI3 Dans Laquelle T Designe Un Metal De Terre Rare Ou Lyttrium. C. R. Heb.
Séances Acad. Sci. Ser. B. 264, 1056–1967
15 Metallic Magnetic Materials 785

163. Gignoux, D., Lemaire, R., Molho, P., Tasset, F.: Onset of magnetism in the yttrium-nickel
compounds: II. Very weak itinerant ferromagnetism in YNi3. J. Magn. Magn. Mater. 21,
307–315 (1980)
164. Lemaire, R., Paccard, D., Pauthenet, R., Schweizer, J.: Magnetic behavior of cobalt and of
nickel in compounds with rare earth metals. J. Appl. Phys. 39, 1092–1093 (1968)
165. Gignoux, D., Lemaire, R., Molho, P.: Onset of magnetism in the yttrium-nickel compounds:
I. Collective electron metamagnetism in Y2Ni17. J. Magn. Magn. Mater. 21, 119–124
(1980)
166. Sadron, S.: Ferromagnetic moments of the elements and the periodic system. Ann. Phvs. 17,
371 (1932)
167. Tazuke, Y., Nakabayashi, R., Murayama, S., Sakakibara, T., Goto, T.: Magnetism of R2Ni7
and RNi3 (R= Y, La, Ce). Phys. B. 186, 596–598 (1993)
168. Buschow, K.: Magnetic properties of La 2 Ni 7 and its hydride. J. Magn. Magn. Mater. 40,
224–226 (1983)
169. Kawamiya, N., Adachi, K.: Magnetic properties of ordered and disordered Ni 1− x Fe x Pt.
Trans. Jpn. Inst. Met. 16, 327–332 (1975)
170. Parra, R., Cable, J.: Neutron study of magnetic-moment distribution in Ni-Pt alloys. Phys.
Rev. B. 21, 5494 (1980)
171. Marian, V.: Ferromagnetic Curie points and the absolute saturation of some nickel alloys.
Ann. Phys. 7, 459–527 (1937)
172. De Boer, F., Schinkel, C., Biesterbos, J., Proost, S.: Exchange-enhanced paramagnetism and
weak ferromagnetism in the Ni3Al and Ni3Ga phases; Giant moment inducement in Fe-
Doped Ni3Ga. J. Appl. Phys. 40, 1049–1055 (1969)
173. Heczko, O., Scheerbaum, N., Gutfleisch, O.: Magnetic shape memory phenomena. In:
Nanoscale Magnetic Materials and Applications, pp. 399–439. Springer (2009)
174. Tickle, R., James, R.: Magnetic and magnetomechanical properties of Ni 2 MnGa.
J. Mag.Mag. Mater. 195, 627–638 (1999)
175. Planes, A., Mañosa, L., Acet, M.: Magnetocaloric effect and its relation to shape-memory
properties in ferromagnetic Heusler alloys. J. Phys.: Condens. Matter. 21, 233201 (2009)
176. Ullakko, K., Huang, J., Kantner, C., O’handley, R., Kokorin, V.: Large magnetic-field-induced
strains in Ni2MnGa single crystals. Appl. Phys. Lett. 69, 1966–1968 (1996)
177. Wu, G., Yu, C., Meng, L., Chen, J., Yang, F., Qi, S., et al.: Giant magnetic-field-induced
strains in Heusler alloy NiMnGa with modified composition. Appl. Phys. Lett. 75, 2990–2992
(1999)
178. Yu, C., Wang, W., Chen, J., Wu, G., Yang, F., Tang, N., et al.: Magnetic-field-induced strains
and magnetic properties of Heusler alloy Ni52Mn23Ga25. J. Appl. Phys. 87, 6292–6294
(2000)
179. Murray, S.J., Marioni, M., Allen, S., O’handley, R., Lograsso, T.A.: 6% magnetic-field-
induced strain by twin-boundary motion in ferromagnetic Ni–Mn–Ga. Appl. Phys. Lett. 77,
886–888 (2000)
180. Likhachev, A., Ullakko, K.: Magnetic-field-controlled twin boundaries motion and giant
magneto-mechanical effects in Ni–Mn–Ga shape memory alloy. Phys. Lett. A. 275, 142–151
(2000)
181. Heczko, O., Sozinov, A., Ullakko, K.: Giant field-induced reversible strain in magnetic shape
memory NiMnGa alloy. IEEE Trans. Magn. 36, 3266–3268 (2000)
182. Wang, W., Wu, G., Chen, J., Gao, S., Zhan, W., Wen, G., et al.: Intermartensitic transformation
and magnetic-field-induced strain in Ni52Mn24. 5Ga23. 5 single crystals. Appl. Phys. Lett.
79, 1148–1150 (2001)
183. Sozinov, A., Likhachev, A., Lanska, N., Ullakko, K.: Giant magnetic-field-induced strain in
NiMnGa seven-layered martensitic phase. Appl. Phys. Lett. 80, 1746–1748 (2002)
184. Liang, T., Jiang, C., Xu, H.: Temperature dependence of transformation strain and magnetic-
field-induced strain in Ni 51 Mn 24 Ga 25 single crystal. Mater. Sci. Eng., A. 402, 5–8 (2005)
185. Jiang, C., Wang, J., Xu, H.: Temperature dependence of the giant magnetostrain in a NiMnGa
magnetic shape memory alloy. Appl. Phys. Lett. 86, 252508 (2005)
786 J. P. Liu et al.

186. Callaway, J., Sehitoglu, H., Hamilton, R., Aslantas, K., Miller, N., Maier, H., et al.: Magnetic
shape memory in Ni2MnGa as influenced by applied stress. Appl. Phys. Lett. 89, 221905
(2006)
187. Popov, A., Belozerov, E., Sagaradze, V., Pecherkina, N., Kabanova, I., Gaviko, V., et al.:
Martensitic transformations and magnetic-field-induced strains in Ni50Mn50− x Gax alloys.
Physi Met Metallogr. 102, 140–148 (2006)
188. Pagounis, E., Chulist, R., Szczerba, M., Laufenberg, M.: Over 7% magnetic field-induced
strain in a Ni-Mn-Ga five-layered martensite. Appl. Phys. Lett. 105, 052405 (2014)
189. Pagounis, E., Chulist, R., Szczerba, M., Laufenberg, M.: High-temperature magnetic shape
memory actuation in a Ni–Mn–Ga single crystal. Scr. Mater. 83, 29–32 (2014)
190. Scheerbaum, N., Hinz, D., Gutfleisch, O., Müller, K.-H., Schultz, L.: Textured polymer
bonded composites with Ni–Mn–Ga magnetic shape memory particles. Acta Mater. 55, 2707–
2713 (2007)
191. Kauffmann-Weiss, S., Scheerbaum, N., Liu, J., Klauss, H., Schultz, L., Mäder, E., et al.:
Reversible magnetic field induced strain in Ni2MnGa-polymer-composites. Adv. Eng. Mater.
14, 20–27 (2012)
192. Scheerbaum, N., Heczko, O., Liu, J., Hinz, D., Schultz, L., Gutfleisch, O.: Magnetic field-
induced twin boundary motion in polycrystalline Ni–Mn–Ga fibres. New J. Phys. 10, 073002
(2008)
193. Boonyongmaneerat, Y., Chmielus, M., Dunand, D.C., Müllner, P.: Increasing magnetoplastic-
ity in polycrystalline Ni-Mn-Ga by reducing internal constraints through porosity. Phys. Rev.
Lett. 99, 247201 (2007)
194. Chmielus, M., Zhang, X., Witherspoon, C., Dunand, D., Müllner, P.: Giant magnetic-field-
induced strains in polycrystalline Ni–Mn–Ga foams. Nat. Mater. 8, 863–866 (2009)
195. Pötschke, M., Weiss, S., Gaitzsch, U., Cong, D., Hürrich, C., Roth, S., et al.: Magnetically
resettable 0.16% free strain in polycrystalline Ni–Mn–Ga plates. Scr. Mater. 63, 383–386
(2010)
196. Gaitzsch, U., Romberg, J., Pötschke, M., Roth, S., Müllner, P.: Stable magnetic-field-induced
strain above 1% in polycrystalline Ni–Mn–Ga. Scr. Mater. 65, 679–682 (2011)
197. Gaitzsch, U., Pötschke, M., Roth, S., Rellinghaus, B., Schultz, L.: A 1% magnetostrain in
polycrystalline 5M Ni–Mn–Ga. Acta Mater. 57, 365–370 (2009)
198. Liang, T., Jiang, C., Xu, H., Liu, Z., Zhang, M., Cui, Y., et al.: Phase transition strain and
large magnetic field induced strain in Ni 50.5 Mn 24 Ga 25.5 unidirectionally solidified alloy.
J. Magn. Magn. Mater. 268, 29–32 (2004)
199. Kohl, M., Agarwal, A., Chernenko, V., Ohtsuka, M., Seemann, K.: Shape memory effect and
magnetostriction in polycrystalline Ni–Mn–Ga thin film microactuators. Mater. Sci. Eng., A.
438, 940–943 (2006)
200. Liu, G., Chen, J., Liu, Z., Dai, X., Wu, G., Zhang, B., et al.: Martensitic transformation and
shape memory effect in a ferromagnetic shape memory alloy: Mn2NiGa. Appl. Phys. Lett.
87, 262504 (2005)
201. Liu, J., Zheng, H., Xia, M., Huang, Y., Li, J.: Martensitic transformation and magnetic
properties in Heusler CoNiGa magnetic shape memory alloys. Scr. Mater. 52, 935–938
(2005)
202. Morito, H., Oikawa, K., Fujita, A., Fukamichi, K., Kainuma, R., Ishida, K.: Enhancement
of magnetic-field-induced strain in Ni–Fe–Ga–Co Heusler alloy. Scr. Mater. 53, 1237–1240
(2005)
203. Kainuma, R., Imano, Y., Ito, W., Sutou, Y., Morito, H., Okamoto, S., et al.: Magnetic-field-
induced shape recovery by reverse phase transformation. Nature. 439, 957–960 (2006)
204. Monroe, J., Karaman, I., Basaran, B., Ito, W., Umetsu, R., Kainuma, R., et al.: Direct mea-
surement of large reversible magnetic-field-induced strain in Ni–Co–Mn–In metamagnetic
shape memory alloys. Acta Mater. 60, 6883–6891 (2012)
205. Morito, H., Fujita, A., Oikawa, K., Ishida, K., Fukamichi, K., Kainuma, R.: Stress-assisted
magnetic-field-induced strain in Ni–Fe–Ga–Co ferromagnetic shape memory alloys. Appl.
Phys. Lett. 90, 062505 (2007)
15 Metallic Magnetic Materials 787

206. Morito, H., Oikawa, K., Fujita, A., Fukamichi, K., Kainuma, R., Ishida, K.: A large magnetic-
field-induced strain in Ni–Fe–Mn–Ga–Co ferromagnetic shape memory alloy. J. Alloys
Compd. 577, S372–S375 (2013)
207. Omori, T., Watanabe, K., Umetsu, R., Kainuma, R., Ishida, K.: Martensitic transformation
and magnetic field-induced strain in Fe–Mn–Ga shape memory alloy. Appl. Phys. Lett. 95,
082508 (2009)
208. Sugiyama, M., Oshima, R., Fujita, F.E.: Martensitic transformation in the Fe–Pd alloy system.
Trans. Jpn. Ins. Met. 25, 585–592 (1984)
209. James, R.D., Wuttig, M.: Magnetostriction of martensite. Philos. Mag. A. 77, 1273–1299
(1998)
210. Kakeshita, T., Fukuda, T.: Giant magnetostriction in Fe3Pt and FePd ferromagnetic shape-
memory alloys. Mater. Sci. For. 394, 531–536 (2002)
211. Sakon, T., Takaha, A., Obara, K., Dejima, K., Nojiri, H., Motokawa, M., et al.: Magnetic-
field-induced strain of shape-memory alloy Fe3Pt studied by a capacitance method in a pulsed
magnetic field. Jpn. J. Appl. Phys. 46, 146 (2007)
212. Fukuda, T., Kakeshita, T.: Giant magnetic field induced strain in ferromagnetic shape memory
alloys and its condition. Mater. Sci. Technol. 24, 890–895 (2008)
213. Liu, E., Zhu, W., Feng, L., Chen, J., Wang, W., Wu, G., et al.: Vacancy-tuned paramagnet-
ic/ferromagnetic martensitic transformation in Mn-poor Mn1-xCoGe alloys. Europhys. Lett.
91, 17003 (2010)
214. Liu, E., Wang, W., Feng, L., Zhu, W., Li, G., Chen, J., et al.: Stable magnetostructural coupling
with tunable magnetoresponsive effects in hexagonal ferromagnets. Nat. Commun. 3, 873
(2012)
215. Zhang, C., Shi, H., Ye, E., Nie, Y., Han, Z., Wang, D.: Magnetostructural transition and
magnetocaloric effect in MnCoGe–NiCoGe system. J. Alloys Compd. 639, 36–39 (2015)
216. Wei, Z.Y., Liu, E.K., Li, Y., Xu, G.Z., Zhang, X.M., Liu, G.D., et al.: Unprecedentedly
wide Curie-temperature windows as phase-transition design platform for tunable magneto-
multifunctional materials. Adv. Electron. Mater. 1 (2015)
217. Fujita, A., Fukamichi, K., Gejima, F., Kainuma, R., Ishida, K.: Magnetic properties and large
magnetic-field-induced strains in off-stoichiometric Ni-Mn-Al Heusler alloys. Appl. Phys.
Lett. 77, 3054 (2000)
218. Liu, Z., Zhang, M., Cui, Y., Zhou, Y., Wang, W., Wu, G., et al.: Martensitic transformation and
shape memory effect in ferromagnetic Heusler alloy Ni2FeGa. Appl. Phys. Lett. 82, 424–426
(2003)
219. Krenke, T., Acet, M., Wassermann, E.F., Moya, X., Mañosa, L., Planes, A.: Ferromagnetism
in the austenitic and martensitic states of Ni− Mn− In alloys. Phys. Rev. B. 73, 174413 (2006)
220. Zhang, M., Brück, E., de Boer, F.R., Wu, G.: Magnetic, martensitic transformation, magne-
tostriction and shape memory effect in Co50Ni20Ga30 melt-spun ribbons. J. Phys. D: Appl.
Phys. 38, 1361 (2005)
221. Li, Z., Xu, K., Yang, H., Zhang, Y., Jing, C.: Magnetostrain and magnetocaloric effect by
field-induced reverse martensitic transformation for Pd-doped Ni45Co5Mn37In13 Heusler
alloy. J. Appl. Phys. 117, 223904 (2015)
222. Jing, C., Wang, X., Liao, P., Li, Z., Yang, Y., Kang, B., et al.: Martensitic phase transition,
inverse magnetocaloric effect, and magnetostrain in Ni50Mn37-xFexIn13 Heusler alloys.
J. Appl. Phys. 114, 063907 (2013)
223. Kainuma, R., Imano, Y., Ito, W., Morito, H., Sutou, Y., Oikawa, K., et al.: Metamagnetic
shape memory effect in a Heusler-type Ni43Co7Mn39Sn11 polycrystalline alloy. Appl. Phys.
Lett. 88, 192513 (2006)
224. Liao, P., Jing, C., Zheng, D., Li, Z., Kang, B., Deng, D., et al.: Tuning martensitic
transformation, large magnetoresistance and strain in Ni 50− x Fe x Mn 36 Sn 14 Heusler
alloys. Solid State Commun. 217, 28–33 (2015)
225. Chernenko, V., Barandiarán, J., L’vov, V., Gutiérrez, J., Lázpita, P., Orue, I.: Temperature
dependent magnetostrains in polycrystalline magnetic shape memory Heusler alloys. J. Alloys
Compd. 577, S305–S308 (2013)
788 J. P. Liu et al.

226. Li, Z., Jing, C., Zhang, H., Qiao, Y., Cao, S., Zhang, J., et al.: A considerable metamagnetic
shape memory effect without any prestrain in Ni46Cu4Mn38Sn12 Heusler alloy. J. Appl.
Phys. 106, 083908 (2009)
227. Zhu, W., Liu, E., Feng, L., Tang, X., Chen, J., Wu, G., et al.: Magnetic-field-induced
transformation in FeMnGa alloys. Appl. Phys. Lett. 95, 222512–222511 (2009)
228. Wei, Z., Liu, E., Chen, J., Li, Y., Liu, G., Luo, H., et al.: Realization of multifunctional shape-
memory ferromagnets in all-d-metal Heusler phases. Appl. Phys. Lett. 107, 022406 (2015)
229. Felser, C., Fecher, G.H., Balke, B.: Spintronics: a challenge for materials science and solid-
state chemistry. Angew. Chem. Int. Ed. 46, 668–699 (2007)
230. Graf, T., Felser, C., Parkin, S.S.: Simple rules for the understanding of Heusler compounds.
Prog. Solid State Chem. 39, 1–50 (2011)
231. Villar, P., Calvert, L.: Pearson’s Handbook of Crystallographic Data for Intermetallic Phases,
2nd edn. ASM International, Materials Park (1996)
232. Kandpal, H.C., Felser, C., Seshadri, R.: Covalent bonding and the nature of band gaps in some
half-Heusler compounds. J. Phys. D: Appl. Phys. 39, 776 (2006)
233. Pan, Y., Nikitin, A., Bay, T., Huang, Y., Paulsen, C., Yan, B., et al.: Superconductivity and
magnetic order in the noncentrosymmetric half-Heusler compound ErPdBi. Europhys. Lett.
104, 27001 (2013)
234. Nikitin, A., Pan, Y., Mao, X., Jehee, R., Araizi, G., Huang, Y., et al.: Magnetic and
superconducting phase diagram of the half-Heusler topological semimetal HoPdBi. J. Phys.:
Condens. Matter. 27, 275701 (2015)
235. De Groot, R., Mueller, F., Van Engen, P., Buschow, K.: New class of materials: half-metallic
ferromagnets. Phys. Rev. Lett. 50, 2024 (1983)
236. Felser, C., Wollmann, L., Chadov, S., Fecher, G.H., Parkin, S.S.: Basics and prospective of
magnetic Heusler compounds. APL Mater. 3, 041518 (2015)
237. Felser, C., Wollmann, L., Chadov, S., Fecher, G.H., Parkin, S.S.P.: Heusler alloys. In: Felser,
C., Hirohata, A. (eds.) Springer Series in Materials Science, vol. 222, pp. 37–48. Springer
International Publishing (2016)
238. Nishino, Y., Kato, M., Asano, S., Soda, K., Hayasaki, M., Mizutani, U.: Semiconductorlike
behavior of electrical resistivity in Heusler-type Fe 2 VAl compound. Phys. Rev. Lett. 79,
1909 (1997)
239. Galanakis, I., Dederichs, P., Papanikolaou, N.: Slater-Pauling behavior and origin of the half-
metallicity of the full-Heusler alloys. Phys. Rev. B. 66, 174429 (2002)
240. Felser, C., Fecher, G.H.: Spintronics: From Materials to Devices. Springer Science &
Business Media (2013)
241. Kandpal, H.C., Ksenofontov, V., Wojcik, M., Seshadri, R., Felser, C.: Electronic structure,
magnetism and disorder in the Heusler compound Co2TiSn. J. Phys. D: Appl. Phys. 40, 1587
(2007)
242. Liu, H.-x., Kawami, T., Moges, K., Uemura, T., Yamamoto, M., Shi, F., et al.: Influence
of film composition in quaternary Heusler alloy Co2 (Mn, Fe) Si thin films on tunnelling
magnetoresistance of Co2 (Mn, Fe) Si/MgO-based magnetic tunnel junctions. J. Phys. D:
Appl. Phys. 48, 164001 (2015)
243. Jourdan, M., Minár, J., Braun, J., Kronenberg, A., Chadov, S., Balke, B., et al.: Direct
observation of half-metallicity in the Heusler compound Co2MnSi. Nat. Commun. 5 (2014)
244. Wurmehl, S., Fecher, G.H., Kandpal, H.C., Ksenofontov, V., Felser, C., Lin, H.-J., et al.:
Geometric, electronic, and magnetic structure of Co 2 FeSi: Curie temperature and magnetic
moment measurements and calculations. Phys. Rev. B. 72, 184434 (2005)
245. Kübler, J., Fecher, G., Felser, C.: Understanding the trend in the Curie temperatures of Co
2-based Heusler compounds: Ab initio calculations. Phys. Rev. B. 76, 024414 (2007)
246. Wurmehl, S., Kandpal, H.C., Fecher, G.H., Felser, C.: Valence electron rules for prediction
of half-metallic compensated-ferrimagnetic behaviour of Heusler compounds with complete
spin polarization. J. Phys.: Condens. Matter. 18, 6171 (2006)
247. Wollmann, L., Chadov, S., Kübler, J., Felser, C.: Magnetism in cubic manganese-rich Heusler
compounds. Phys. Rev. B. 90, 214420 (2014)
15 Metallic Magnetic Materials 789

248. Galanakis, I., Özdoǧan, K., Aktaş, B., Şaşinoǧlu, E.: Ferrimagnetism and antiferro-
magnetism in half-metallic Heusler alloys. Phys. Status Solidi A. 205, 1036–1039
(2008)
249. Wollmann, L., Chadov, S., Kübler, J., Felser, C.: Magnetism in tetragonal manganese-rich
Heusler compounds. Phys. Rev. B. 92, 064417 (2015)
250. Kübler, J., William, A., Sommers, C.: Formation and coupling of magnetic moments in
Heusler alloys. Phys. Rev. B. 28, 1745 (1983)
251. Ouardi, S., Fecher, G.H., Stinshoff, R., Felser, C., Kubota, T., Mizukami, S., et al.:
Stoichiometry dependent phase transition in Mn-Co-Ga-based thin films: from cubic in-plane,
soft magnetized to tetragonal perpendicular, hard magnetized. Appl. Phys. Lett. 101 (2012)
252. Balke, B., Fecher, G.H., Winterlik, J., Felser, C.: Mn3Ga, a compensated ferrimagnet with
high Curie temperature and low magnetic moment for spin torque transfer applications. Appl.
Phys. Lett. 90, 2504 (2007)
253. Coey, J.: Permanent magnets: plugging the gap. Scr. Mater. 67, 524–529 (2012)
254. Winterlik, J., Chadov, S., Gupta, A., Alijani, V., Gasi, T., Filsinger, K., et al.: Design scheme of
new tetragonal Heusler compounds for spin-transfer torque applications and its experimental
realization. Adv. Mater. 24, 6283–6287 (2012)
255. Siewert, M., Gruner, M., Dannenberg, A., Chakrabarti, A., Herper, H., Wuttig, M., et al.:
Designing shape-memory Heusler alloys from first-principles. Appl. Phys. Lett. 99, 191904
(2011)
256. Suits, J.C.: New magnetic compounds with Heusler and Heusler-related structures. Phys. Rev.
B. 14, 4131 (1976)
257. Kurt, H., Rode, K., Stamenov, P., Venkatesan, M., Lau, Y.-C., Fonda, E., et al.: Cubic Mn 2
Ga thin films: crossing the spin gap with ruthenium. Phys. Rev. Lett. 112, 027201 (2014)
258. Galanakis, I., Şaşıoğlu, E.: High TC half-metallic fully-compensated ferrimagnetic Heusler
compounds. Appl. Phys. Lett. 99, 052509 (2011)
259. Nayak, A.K., Nicklas, M., Chadov, S., Khuntia, P., Shekhar, C., Kalache, A., et al.: Design
of compensated ferrimagnetic Heusler alloys for giant tunable exchange bias. Nat. Mater. 14,
679–684 (2015)
260. Yoon, S., Booth, J.: Structural and magnetic properties of Fe 3− x Mn x Si alloys. Phys. Lett.
A. 48, 381–382 (1974)
261. Mohn, P., Supanetz, E.: Spin ordering in Fe3-x Mn x Si Heusler alloys. Philos. Mag. B. 78,
629–636 (1998)
262. Meshcheriakova, O., Chadov, S., Nayak, A., Rößler, U., Kübler, J., André, G., et al.: Large
noncollinearity and spin reorientation in the novel Mn 2 RhSn Heusler magnet. Phys. Rev.
Lett. 113, 087203 (2014)
263. Buschow, K.: The crystal structure of Th2Co7. Acta Crystallogr. Sect. B. 26, 1389–1392
(1970)
264. Buschow, K.: Structural and magnetic characteristics of Th–Co and Th–Fe compounds.
J. Appl. Phys. 42, 3433–3437 (1971)
265. Bartashevich, M., Deryagin, A., Kudrevatykh, N., Tarasov, E.: High-temperature metam-
agnetism of the Y2CO7H6 hydride. Zhurnal Eksperimentalnoi I Teoreticheskoi Fiziki. 84,
1140–1144 (1983)
266. Andreev, A., Bartashevich, M.: AV Der3’agin, tnd Ye, N. Tarasov. Phys. Met. Metallogr. 62,
66 (1986)
267. Kuz’min, M., Skokov, K., Radulov, I., Schwöbel, C., Foro, S., Donner, W., et al.: Magnetic
anisotropy of La2Co7. J. Appl. Phys. 118, 053905 (2015)
268. Buschow, K.: Magnetic properties of CeCo3, Ce2Co7 and CeNi3 and their ternary hydrides.
J. Less-Common Met. 72, 257–263 (1980)
269. Buschow, K.H.J.: Colloq. Int. CNRS, Les Elements de Terres Rares 1969, 1, 101 (1970)
270. Ray, A.E., Biermann, A.T., Hammer, R.S., Davison, J.E.: Cobalt. 4, 103 (1973)
271. Bartashevich, M., Andreev, A., Tarasov, E., Goto, T., Yamaguchi, M.: Magnetic properties
and spontaneous magnetostriction of a Sm2Co7 single crystal. Phys. B. 183, 369–378
(1993)
790 J. P. Liu et al.

272. Bartashevich, M., Goto, T., Yamaguchi, M.: Field induced magnetic phase transition and
magnetostriction in ErCo 3, HoCo 3 and Nd 2 Co 7 single crystals. J. Magn. Magn. Mater.
111, 83–89 (1992)
273. Lemaire, R.: Magnetic properties of the intermetallic compounds of cobalt with the rare earth
metals and yttrium. Cobalt, 201–211 (1966)
274. Andreev, A., Tarasov, E., Deryagin, A., Zadvorkin, S.: Influence of symmetry of the
crystalline structure on the magnetic anisotropy of Tb2Co7. Phys. Status Solidi A. 71, K245–
K247 (1982)
275. Ostertag, W.: The crystal structure of Er2Co7 and other rare earth-cobalt compounds
R2Co7 (R = Gd, Tb, Dy, Ho, Tm, Lu, Y). J. Less-Common Met. 13, 385–390
(1967)
276. Andreev, A., Zadvorkin, S., Tarasov, E.: Thermal expansion anomalies and spontaneous
magnetostriction in Tm 2 Co 7. J. Alloys Compd. 189, 187–190 (1992)
277. Parker, F., Oesterreicher, H.: Magnetic properties of La2Ni7. J. Less-Common Met. 90, 127–
136 (1983)
278. Lemaire, R., Paccard, D., Pauthenet, R.: Structural and – magnetic properties of alloys of
composition T 2 NI 7, where T is a rare-earth metal or yttrium. Compt. Rend. 265(23), 1280–
1282 (1967)
279. Virkar, A.V., Raman, A.: Crystal structures of AB3 and A2B7 rare earth-nickel phases.
J. Less-Common Met. 18, 59–66 (1969)
280. Buschow, K., Van Der Goot, A.: The crystal structure of rare-earth nickel compounds of the
type R 2 Ni 7. J. Less-Common Met. 22, 419–428 (1970)
281. Bhattacharyya, A., Giri, S., Majumdar, S.: Successive magnetic transitions and low tempera-
ture magnetocaloric effect in RE2Ni7 (RE=Dy, Ho). J. Magn. Magn. Mater. 323, 1484–1489
(2011)
282. Tsvyashchenko, A.V.: High-pressure synthesis of YB3CO, YB3NI compounds. J. Less-
Common Met. 118, 103–107 (1986)
283. Palenzona, A., Cirafici, S.: The equilibrium diagram of the Th-Ni system. J. Less-Common
Met. 142, 311–317 (1988)
284. Schneider, G., Landgraf, F., Villas-Boas, V., Bezerra, G., Missell, F., Ray, A.: New stable
phase in the binary Fe-Nd system. Mater. Lett. 8, 472–476 (1989)
285. Moreau, J.M., Paccard, L., Nozieres, J.P., Missell, F.P., Schneider, G., Villas-Boas, V.: A new
phase in the Nd-Fe system: Crystal structure of Nd5Fe17. J. Less Common Met. 163, 245–251
(1990)
286. Nozieres, J., Rechenberg, H.: Magnetic properties of Nd5Fe17. Solid State Commun. 79,
21–24 (1991)
287. Stadelmaier, H., Schneider, G., Henig, E.-T., Ellner, M.: Magnetic Fe17R5 in the Fe-Nd and
Fe (-Ti)-Sm systems, and other phases in Fe-Nd. Mater. Lett. 10, 303–309 (1991)
288. Maruyama, F., Amako, Y.: Magnetic properties of Sm 5 (Fe 1− x T x) 17 (T= Ti and V)
compounds. J. Alloys Compd. 474, 1–3 (2009)
289. Saito, T.: High coercivity in Sm 5 Fe 17 melt-spun ribbon. J. Alloys Compd. 440, 315–318
(2007)
290. Saito, T., Ichihara, M.: Synthesis and magnetic properties of Sm 5 Fe 17 hard magnetic phase.
Scr. Mater. 57, 457–460 (2007)
291. Yelon, W., Luo, H., Chen, M., Missell, F.: Magnetic and crystallographic structure of
Nd5Fe17. J. Appl. Phys. 85, 5693–5695 (1999)
292. Chu, Z., Yelon, W., Missell, F., Murakami, R.: Site occupancy of Sm in (Nd1− xSmx) 5
(Fe1− yTiy) 17. J. Appl. Phys. 87, 6704–6706 (2000)
293. Saito, T.: Annealing of amorphous Sm5Fe17 melt-spun ribbon. Mater. Trans. 49, 1446–1450
(2008)
294. Stadelmaier, H., Schneider, G., Ellner, M.: A CaCu5-type iron-neodymium phase stabilized
by rapid solidification. J. Less-Common Met. 115, L11–L14 (1986)
295. Moze, O., Pareti, L., Paoluzi, A., Buschow, K.: Magnetic structure and anisotropy of Ga-and
Al-substituted LaCo 5 and YCo 5 intermetallics. Phys. Rev. B. 53, 11550 (1996)
15 Metallic Magnetic Materials 791

296. Alameda, J., Givord, D., Lemaire, R., Lu, Q.: Co energy and magnetization anisotropies in
RCo5 intermetallics between 4.2 K and 300 K. J. Appl. Phys. 52, 2079–2081 (1981)
297. Alameda, J., Deportes, J., Givord, D., Lemaire, R., Lu, Q.: Large magnetization anisotropy in
uniaxial YCo5 intermetallic. J. Magn. Magn. Mater. 15, 1257–1258 (1980)
298. Barbara, B., Uehara, M.: Anisotropy and coercivity in SmCo 5-based compounds. IEEE
Trans. Magn. 12, 997–999 (1976)
299. Gubbens, P., Van der Kraan, A.: Magnetic properties of ThFe 5. J. Magn. Magn. Mater. 9,
349–354 (1978)
300. Cadieu, F., Cheung, T., Wickramasekara, L., Aly, S.: Magnetic properties of a metastable
Sm-Fe phase synthesized by selectively thermalized sputtering. J. Appl. Phys. 55, 2611–2613
(1984)
301. Alameda, J., Givord, D., Lemaire, R., Lu, Q., Palmer, S., Tasset, F.: Reduced 4f-moment of
the Nd ground state in NdCo5. Le Journal de Physique Colloques. 43, C7-133–C7-139 (1982)
302. Schweizer, J., Tasset, F.: Polarised neutron study of the RCo5 intermetallic compounds. I. The
cobalt magnetisation in YCo5. J. Phys. F: Met. Phys. 10, 2799 (1980)
303. Decrop, B., Deportes, J., Lemaire, R.: Magnetic structure of HoCo 5 below room temperature.
J. Less-Common Met. 94, 199–203 (1983)
304. Coroian, N., Klosek, V., Isnard, O.: The influence of substituting Si for Co on the magnetic
properties of PrCo 5. J. Alloys Compd. 427, 5–10 (2007)
305. Barthem, V., Gignoux, D., Nait-Saada, A., Schmitt, D., Creuzet, G.: Magnetic and magnetoe-
lastic properties of PrNi 5 single crystal. Phys. Rev. B. 37, 1733 (1988)
306. Coldea, M., Andreica, D., Bitu, M., Crisan, V.: Spin fluctuations in YNi 5 and CeNi 5.
J. Magn. Magn. Mater. 157, 627–628 (1996)
307. Grechnev, G., Logosha, A., Panfilov, A., Kuchin, A., Vasijev, A.: Effect of pressure on the
magnetic properties of YNi5, LaNi5, and CeNi5. Low Temp. Phys. 37, 138 (2011)
308. Burzo, E., Ursu, I.: Paramagnetic resonance and magnetic measurements on GdNi 5
compound. Solid State Commun. 9, 2289–2292 (1971)
309. Barthem, V., Gignoux, D., Naitsaada, A., Schmitt, D., Takeuchi, A.Y.: Magnetic-properties of
the hexagonal NDNI5 and NDCU5 Compounds. J. Magn. Magn. Mater. 80, 142–148 (1989)
310. Gubbens, P., Van Der Kraan, A., Buschow, K.: Crystal-field effects and magnetic behavior in
R Ni 5 and R Co 5+ x rare-earth compounds. Phys. Rev. B. 39, 12548 (1989)
311. Gignoux, D., Nait-Saada, A., De La Bâthie, R.P.: Magnetic properties of TbNi5 and HoNi5
single crystals. Le Journal de Physique Colloques. 40, C5-188–C5-190 (1979)
312. Sada, A.N.: Ph.D. thesis, University of Grenoble (1980)
313. Givord, D., Laforest, J., Schweizer, J., Tasset, F.: Temperature dependence of the samarium
magnetic form factor in SmCo5. J. Appl. Phys. 50, 2008–2010 (1979)
314. Buschow, K., Van Diepen, A., De Wijn, H.: Crystal-field anisotropy of Sm 3+ in SmCo 5.
Solid State Commun. 15, 903–906 (1974)
315. Laforest, J.: Ph.D. thesis, Université de Grenoble (1981)
316. Sankar, S., Rao, V., Segal, E., Wallace, W., Frederick, W., Garrett, H.: Magnetocrystalline
anisotropy of Sm Co 5 and its interpretation on a crystal-field model. Phys. Rev. B. 11, 435
(1975)
317. Von Ranke, P., Mota, M., Grangeia, D., Carvalho, A.M.G., Gandra, F., Coelho, A., et al.:
Magnetocaloric effect in the R Ni 5 (R= Pr, Nd, Gd, Tb, Dy, Ho, Er) series. Phys. Rev. B. 70,
134428 (2004)
318. Skokov, K.P., Pastushenkov, Y.G., Koshkid’ko, Y.S., Shütz, G., Goll, D., Ivanova, T.I., et al.:
Magnetocaloric effect, magnetic domain structure and spin-reorientation transitions in HoCo5
single crystals. J. Magn. Magn. Mater. 323, 447–450 (2011)
319. Givord, D., Laforest, J., Lemaire, R.: Polarized neutron study of the itinerant electron
metamagnetism in ThCo5. J. Appl. Phys. 50, 7489–7491 (1979)
320. Givord, D., Laforest, J., Lemaire, R.: Magnetic transition in ThCo5 due to change of Co-
moment. Phys. B. 86–88, Part 1, 204–206 (1977)
321. Ballou, R., Shimizu, M., Voiron, J.: Pressure effects on the metamagnetic transition of ThCo
5. J. Magn. Magn. Mater. 84, 23–28 (1990)
792 J. P. Liu et al.

322. Velge, W., Buschow, K.: Magnetic and crystallographic properties of some rare earth cobalt
compounds with CaZn5 structure. J. Appl. Phys. 39, 1717–1720 (1968)
323. Wernick, J., Geller, S.: Transition element–rare earth compounds with Cu5Ca structure. Acta
Crystallogr. 12, 662–665 (1959)
324. Klein, H.P., Menth, A., Perkins, R.S.: Magnetocrystalline anisotropy of light rare-earth cobalt
compounds. Phys. B. 80, 153–163 (1975)
325. Ibarra, M., Morellon, L., Algarabel, P., Moze, O.: Single-ion competing magnetic anisotropies
in Pr x Nd 1− x Co 5 intermetallic compounds. Phys. Rev. B. 44, 9368 (1991)
326. Tatsumoto, E., Okamoto, T., Fujii, H., Inoue, C.: Saturation magnetic moment and crystalline
anisotropy of single crystals of light rare earth cobalt compounds RCo5. Le Journal de
Physique Colloques. 32, C1-550–C1-551 (1971)
327. Decrop, B.: Ph.D. thesis, INPG Grenoble (1982)
328. Lemaire, R., Schweizer, J.: Structures magnétiques des composés intermétalliques CeCo5 et
TbCo5. J. Phys. 28, 216–220 (1967)
329. Ohkoshi, M., Kobayashi, H., Katayama, T., Hirano, M., Tsushima, T.: Spin reorientation in
DyCo 5. Phys. B. 86, 195–196 (1977)
330. Buschow, K., Velge, W.: Permanent magnetic materials of rare earth – cobalt compounds.
Philips Research Lab. and Metallurgical Lab, Eindhoven (1969)
331. Givord, D., Laforest, J., Lemaire, R., Lu, Q.: Cobalt magnetism in RCo5-intermetallics: onset
of 3d magnetism and magnetocrystalline anisotropy (r=rare earth or Th). J. Magn. Magn.
Mater. 31–34, Part 1, 191–196 (1983)
332. Van Diepen, A., Buschow, K., Van Wieringen, J.: Study of the Mössbauer effect, magneti-
zation, and crystal structures of the pseudobinary compounds ThCo5− 5xFe5x and ThNi5−
5xFe5x. J. Appl. Phys. 43, 645–650 (1972)
333. Ballou, R., Barthem, V., Gignoux, D.: Crystal field effects in the hexagonal SmNi5 compound.
Phys. B. 149, 340–344 (1988)
334. Oliver, F., West, K., Cohen, R., Buschow, K.: Mossbauer effect of 151Eu in EuNi5, EuMg2
and their hydrides. J. Phys. F: Met. Phys. 8, 701 (1978)
335. De Jesus, V., Oliveira, I., Riedi, P., Guimaraes, A.: 155,157 Gd NMR study of Gd–Ni
intermetallic compounds. J. Magn. Magn. Mater. 212, 125–137 (2000)
336. Mulder, F., Thiel, R., Buschow, K.: 155 Gd Mössbauer effect and magnetic properties of
aluminium-and gallium-substituted GdCu 5 and GdNi 5. J. Alloys Compd. 190, 77–82 (1992)
337. Haldar, A., Dhiman, I., Das, A., Suresh, K., Nigam, A.: Magnetic, magnetocaloric and neutron
diffraction studies on TbNi 5− x M x (M= Co and Fe) compounds. J. Alloys Compd. 509,
3760–3765 (2011)
338. Mulders, A., Kaiser, C., Harker, S., Gubbens, P., Amato, A., Gygax, F., et al.: Muon location
and muon dynamics in DyNi 5. Phys. Rev. B. 67, 014303 (2003)
339. Barthem, V., Gignoux, D., Schmitt, D., Creuzet, G.: Magnetic and magnetoelastic properties
of the hexagonal TmNi 5 compound. J. Magn. Magn. Mater. 78, 56–66 (1989)
340. Zhang, F., Gignoux, D., Schmitt, D., Franse, J., Kayzel, F., Kim-Ngan, N., et al.: Crystalline
electric field and high field magnetization in ErNi 5 single crystal. J. Magn. Magn. Mater.
130, 108–114 (1994)
341. Hodges, J., Bonville, P., Ocio, M.: Magnetic properties of YbNi5 from 170Yb Mössbauer and
magnetisation measurements. Eur. Phys. J. B. 57, 365–370 (2007)
342. Knyazev, Y.V., Lukoyanov, A., Kuz’min, Y.I., Kuchin, A.: Influence of copper impurities on
the evolution of the electronic structure and optical spectra of the LuNi5 compound. Phys.
Solid State. 57, 866–870 (2015)
343. Florio, J.V., Baenziger, N., Rundle, R.: Compounds of thorium with transition metals. II.
Systems with iron, cobalt and nickel. Acta Crystallogr. 9, 367–372 (1956)
344. Narasimhan, K., Do-Dinh, C., Wallace, W., Hutchens, R.: Magnetic properties of the ThCo5−
xNix system. J. Appl. Phys. 46, 4961–4964 (1975)
345. Aubert, G., Gignoux, D., Hennion, B., Michelutti, B., Saada, A.N.: Bulk magnetization study
of a DyNi 5 single crystal. Solid State Commun. 37, 741–743 (1981)
15 Metallic Magnetic Materials 793

346. Boucherle, J., Givord, D., Laforest, J., Schweizer, J., Tasset, F.: Determination of exchange
and crystal field effects in Sm alloys by polarized neutron diffraction. Le Journal de Physique
Colloques. 40, C5-180–C5-182 (1979)
347. Lu, Q.: PhD Thesis, Université de Grenoble (1981).
348. Ibarra, M., Morellon, L., Algarabel, P., Moze, O.: A determination of the crystal electric field
and exchange parameters of Pr 3+ and Nd 3+ ions in RCo 5 intermetallics. J. Magn. Magn.
Mater. 104, 1149–1151 (1992)
349. Decrop, B., Deportes, J., Givord, D., Lemaire, R., Chapert, J.: Study of the magnetization
reorientation in HoCo5. J. Appl. Phys. 53, 1953–1955 (1982)
350. Kren, E., Schweizer, J., Tasset, F.: Polarized-neutron-diffraction study of magnetic moments
in yttrium-cobalt alloys. Phys. Rev. 186, 479 (1969)
351. Buschow, K.: The crystal structures of the rare-earth compounds of the form R 2 Ni 17, R 2
Co 17 and R 2 Fe 17. J. Less-Common Met. 11, 204–208 (1966)
352. Givord, D., Givord, F., Lemaire, R., James, W.J., Shah, J.: Evidence of disordered substitu-
tions in the “Th 2 Ni 17-type” structure. Exact structure determination of the Th-Ni, Y-Ni and
Er-Co compounds. J. Less-Common Met. 29, 389–396 (1972)
353. Isnard, O., Hautot, D., Long, G.J., Grandjean, F.: A structural, magnetic, and Mössbauer
spectral study of Dy2Fe17 and its hydrides. J. Appl. Phys. 88, 2750–2759 (2000)
354. Christensen, A.N., Hasell, R.G.: Acta Chem. Scand A. 34, 6 (1980)
355. Averbuch-Pouchot, M., Chevalier, R., Deportes, J., Kebe, B., Lemaire, R.: Anisotropy of the
magnetization and of the iron hyperfine field in R 2 Fe 17 compounds. J. Magn. Magn. Mater.
68, 190–196 (1987)
356. Burzo, E., Chelkowski, A., Kirchmayr, H.: Magnetic Properties of Metals: Compounds
Between Rare Earth Elements and 3d, 4d or 5d Elements HPJ Wijn, Landolt-Börnstein, New
Series, Group III, vol. 19, (1990)
357. Givord, D., Lemaire, R.: Magnetic transition and anomalous thermal expansion in R 2 Fe 17
compounds. IEEE Trans. Magn. 10, 109–113 (1974)
358. Givord, D., Lemaire, R.: Ferromagnetic-Helimagnetic Transition in the Compounds LuFe 9.
5 and Ce 2 Fe 17. Compt. Rend. 274, 1166–1169 (1972)
359. Prokhnenko, O., Ritter, C., Arnold, Z., Isnard, O., Kamarád, J., Pirogov, A., et al.: Neutron
diffraction studies of the magnetic phase transitions in Ce2Fe17 compound under pressure.
J. Appl. Phys. 92, 385–391 (2002)
360. Kamarád, J., Prokhnenko, O., Prokeš, K., Arnold, Z.: Magnetization and neutron diffraction
studies of Lu2Fe17 under high pressure. J. Phys.: Condens. Matter. 17, S3069 (2005)
361. Kreyssig, A., Chang, S., Janssen, Y., Kim, J., Nandi, S., Yan, J., et al.: Crystallographic phase
transition within the magnetically ordered state of Ce 2 Fe 17. Phys. Rev. B. 76, 054421
(2007)
362. Janssen, Y., Chang, S., Kreyssig, A., Kracher, A., Mozharivskyj, Y., Misra, S., et al.: Magnetic
phase diagram of Ce 2 Fe 17. Phys. Rev. B. 76, 054420 (2007)
363. Prokhnenko, O., Kamarád, J., Prokeš, K., Arnold, Z., Andreev, A.: Helimagnetism of Fe: High
Pressure Study of an Y 2 Fe 17 Single Crystal. Phys. Rev. Lett. 94, 107201 (2005)
364. Kamarád, J., Prokhnenko, O., Prokeš, K., Arnold, Z., Andreev, A.: Pressure induced
helimagnetism in Fe-based (Y 2 Fe 17, Lu 2 Fe 17) intermetallic compounds. J. Magn. Magn.
Mater. 310, 1801–1803 (2007)
365. Prokhnenko, O., Ritter, C., Medvedeva, I., Arnold, Z., Kamarád, J., Kuchin, A.: Neutron
diffraction study of Lu 2 Fe 17 under high pressure. J. Magn. Magn. Mater. 258, 564–566
(2003)
366. Burzo, E.: F. Givord CR hebd. séanc. Acad. Sci. Paris B. 271, 1159 (1970)
367. Elemans, J., Buschow, K.: The Magnetic Structure of Tm2Fe17. Phys. Status Solidi A. 24,
K125–K127 (1974)
368. Isnard, O., Andreev, A., Kuz’min, M., Skourski, Y., Gorbunov, D., Wosnitza, J., et al.: High
magnetic field study of the Tm 2 Fe 17 and Tm 2 Fe 17 D 3.2 compounds. Phys. Rev. B. 88,
174406 (2013)
794 J. P. Liu et al.

369. Deportes, J., Kebe, B., Lemaire, R.: Hyperfine field anisotropy in RE-Fe compounds. J. Magn.
Magn. Mater. 54, 1089–1090 (1986)
370. Koyama, K., Fujii, H., Canfield, P.: Magnetocrystalline anisotropy of a Nd 2 Fe 17 single
crystal. Phys. B. 226, 363–369 (1996)
371. Clausen, K., Nielsen, O.V.: Magnetic anisotropy in single crystals of Ho2Co17 and Ho2Fe17.
J. Magn. Magn. Mater. 23, 237–240 (1981)
372. Isnard, O., Miraglia, S., Guillot, M., Fruchart, D.: High field magnetization measurements
of Sm2Fe17, Sm2Fe17N3, Sm2Fe17D5, and Pr2Fe17, Pr2Fe17N3. J. Appl. Phys. 75, 5988–
5993 (1994)
373. Kou, X., De Boer, F., Grössinger, R., Wiesinger, G., Suzuki, H., Kitazawa, H., et al.: Magnetic
anisotropy and magnetic phase transitions in R 2 Fe 17 with R= Y, Ce, Pr, Nd, Sm, Gd, Tb,
Dy, Ho, Er, Tm and Lu. J. Magn. Magn. Mater. 177, 1002–1007 (1998)
374. Isnard, O., Fruchart, D.: Magnetism in Fe-based intermetallics: relationships between local
environments and local magnetic moments. J. Alloys Compd. 205, 1–15 (1994)
375. Long, G.J., Isnard, O., Grandjean, F.: A Mössbauer spectral study of the magnetic properties
of Ho2Fe17 and Ho2Fe17D3. 8. J. Appl. Phys. 91, 1423–1430 (2002)
376. Schweizer, J., Tasset, F.: Proceedings of the International Conference on Magnetism, 1973
(1974)
377. Ohashi, K.: Present and future of Sm 2 Co 17 magnets. J. Jpn. Inst. Met. (2012)
378. Chin, G.: New magnetic alloys. Science. 208, 888–894 (1980)
379. Chin, G., Wernick, J.: Magnetic Materials, Bulk. In: Encyclopedia of Chemical Technology,
vol. 14, pp. 646–686. Wiley, New York (1981)
380. Strnat, K.: In: Wohlfarth, E.P., Buschow, K.H.J. (eds.) Ferromagnetic Materials, vol. 4.
Elsevier Science Publishers BV (1988)
381. Narasimhan, K., Wallace, W., Hutchens, R., Greedan, J.: Magnetic anisotropy of R2Co17
compounds (R= Er, Tm, Yb). In: American Institute of Physics Conference Series, pp. 1212–
1216 (1974)
382. Deryagin, A., Kudrevatykh, N.: Magnetic anisotropy of single crystals of intermetallic
R2Co17 (R= Tb, Dy, Ho, Lu) compounds. Phys. Status Solidi A. 30, K129–K133 (1975)
383. Deryagin, A., Kudrevatykh, N., Baskhov, Y.: Proceedings of the International Conference on
Magnetism, 1973 (1974)
384. Tereshina, E., Andreev, A.: Crystal structure and magnetic properties of Lu 2 Co 17− x Six
single crystals. Intermetallics. 18, 641–648 (2010)
385. Moze, O., Caciuffo, R., Gillon, B., Kayzel, F.: Magnetization density in Er 2 Co 17. J. Magn.
Magn. Mater. 104, 1394–1396 (1992)
386. Le Caer, G., Malaman, B., Isnard, O., Soubeyroux, J., Fruchart, D., Jacobs, T., et al.:
Magnetic characterisation of the ternary carbide ThFe11C x (1.5≤ x≤ 2) by57Fe Mössbauer
spectroscopy. Hyperfine Interact. 77, 221–234 (1993)
387. Brouha, M., Buschow, K.: Pressure dependence of the Curie temperature of intermetallic
compounds of iron and rare-earth elements, Th, and Zr. J. Appl. Phys. 44, 1813–1816 (1973)
388. Hautot, D., Long, G.J., Grandjean, F., Isnard, O.: Mössbauer spectral study of the magnetic
properties of Ce 2 Fe 17 H x (x= 0, 1, 2, 3, 4, and 5). Phys. Rev. B. 62, 11731 (2000)
389. Dan’kov, S.Y., Ivtchenko, V., Tishin, A., Gschneidner Jr., K., Pecharsky, V.: Magnetocaloric
effect in GdAl2 and Nd2Fe17. In: Advances in Cryogenic Engineering Materials, pp. 397–
404. Springer (2000)
390. Chen, H., Zhang, Y., Han, J., Du, H., Wang, C., Yang, Y.: Magnetocaloric effect in R 2 Fe 17
(R= Sm, Gd, Tb, Dy, Er). J. Magn. Magn. Mater. 320, 1382–1384 (2008)
391. Isnard, O., Miraglia, S., Fruchart, D., Guillot, M.: High field magnetization study of the Gd 2
Fe 17 H x system. IEEE Trans. Magn. 30, 4969–4971 (1994)
392. Kuz’min, M., Skourski, Y., Skokov, K., Müller, K.-H., Gutfleisch, O.: Determining anisotropy
constants from a first-order magnetization process in Tb 2 Fe 17. Phys. Rev. B. 77, 132411
(2008)
393. Zhao, T., Lee, T., Pang, K., Lee, J.: High-field magnetization processes in Tb 2 Fe 17 and Er
2 Fe 17. J. Magn. Magn. Mater. 140, 1009–1010 (1995)
15 Metallic Magnetic Materials 795

394. Nikitin, S., Tereshina, I., Tereshina, E., Suski, W., Drulis, H.: The effect of hydrogen on the
magnetocrystalline anisotropy of R 2 Fe 17 and R (Fe, Ti) 12 (R= Dy, Lu) compounds.
J. Alloys Compd. 451, 477–480 (2008)
395. Franse, J., Radwanski, R., Buschow, K.: Handbook of Magnetic Materials, vol. 7, p. 307.
North-Holland, Amsterdam (1993)
396. Gubbens, P.C.M.: Ph.D. thesis, Delft University Press (1977)
397. Andreev, A., Deryagin, A., Zadvorkin, S., Kudrevatykh, N., Moskalev, V., Levitin, R., et al.:
Physics of Magnetic Materials, p. 21. Kalinin University, Kalinin, USSR (1985)
398. Buschow, K.: Note on the structure and occurrence of ytterbium transition metal compounds.
J. Less-Common Met. 26, 329–333 (1972)
399. Tishin, A.M., Spichkin, Y.I.: The Magnetocaloric Effect and Its Applications. CRC Press
(2003)
400. Bouchet, G., Laforest, J., Lemaire, R., Schweizer, J.: Comptes Rendus Hebdomadaires des
Seances de l’Academie des Sciences, Serie B: Sci. Phys. (1966)
401. Khan, Y.: On the crystal structures of the R2Co17 intermetallic compounds. Acta Crystallogr.
Sect. B. 29, 2502–2507 (1973)
402. Hirosawa, S., Wallace, W.: Effect of Substitution of Zr and Pr on magnetic properties of R 2
Co 17 (R= Er, Yb). J. Magn. Magn. Mater. 30, 238–242 (1982)
403. Merches, M., Wallace, W., Craig, R.: Magnetic and structural characteristics of some 2: 17
rare earth-cobalt systems. J. Magn. Magn. Mater. 24, 97–105 (1981)
404. Herbst, J., Croat, J., Yelon, W.: Structural and magnetic properties of Nd2Fe14B. J. Appl.
Phys. 57, 4086–4090 (1985)
405. Shoemaker, C.B., Shoemaker, D., Fruchart, R.: The structure of a new magnetic phase related
to the sigma phase: iron neodymium boride Nd2Fe14B. Acta Crystallogr. Sect. C: Cryst.
Struct. Commun. 40, 1665–1668 (1984)
406. Givord, D., Li, H., Moreau, J.: Magnetic properties and crystal structure of Nd 2 Fe 14 B.
Solid State Commun. 50, 497–499 (1984)
407. Coey, J.M.D.: In: Coey, J.M.D. (ed.) Chap. 1 in Rare-Earth Iron Permanent Magnets. Oxford
Science Publishing, Oxford (1996)
408. Herbst, J.: R 2 Fe 14 B materials: intrinsic properties and technological aspects. Rev. Mod.
Phys. 63, 819 (1991)
409. Burzo, E.: Permanent magnets based on R-Fe-B and R-Fe-C alloys. Rep. Prog. Phys. 61, 1099
(1998)
410. Franse, J.J.M., Radwanski, R.J.: In: Coey, J.M.D. (ed.) Chap. 2 in Rare-Earth Iron Permanent
Magnets. Oxford Science Publishing, Oxford (1996)
411. Franse, J.J.M., Radwanski, R.J.: In: Buschow, K.H.J. (ed.) Chap. 5 in Handbook of Magnetic
Materials, vol. 7, p. 307. Elsevier B.V (1993) ISBN: 978-0-444-89853-1.
412. Givord, D., Li, H., de La Bathie, R.P.: Magnetic properties of Y 2 Fe 14 B and Nd 2 Fe 14 B
single crystals. Solid State Commun. 51, 857–860 (1984)
413. Fruchart, R., l’Héritier, P., De Reotier, P.D., Fruchart, D., Wolfers, P., Coey, J., et al.:
Mossbauer spectroscopy of R2Fe14B. J. Phys. F: Met. Phys. 17, 483 (1987)
414. Isnard, O., Yelon, W., Miraglia, S., Fruchart, D.: Neutron-diffraction study of the insertion
scheme of hydrogen in Nd2Fe14B. J. Appl. Phys. 78, 1892–1898 (1995)
415. Givord, D., Li, H., Tasset, F.: Polarized neutron study of the compounds Y2Fe14B and
Nd2Fe14B. J. Appl. Phys. 57, 4100–4102 (1985)
416. Wolfers, P., Obbade, S., Fruchart, D., Verhoef, R.: Precise crystal and magnetic structure
determinations. Part I: a neutron diffraction study of Nd 2 Fe 14 B at 20 K. J. Alloys Compd.
242, 74–79 (1996)
417. Wolfers, P., Bacmann, M., Fruchart, D.: Single crystal neutron diffraction investigations of
the crystal and magnetic structures of R 2 Fe 14 B (R= Y, Nd, Ho, Er). J. Alloys Compd. 317,
39–43 (2001)
418. Obbade, S., Wolfers, P., Fruchart, D., Argoud, R., Muller, J., Palacios, E.: A precise crystal
structure determination. Part II: an X-ray four-circle study of Nd 2 Fe 14 B at 20 and 290 K.
J. Alloys Compd. 242, 80–84 (1996)
796 J. P. Liu et al.

419. Herbst, J., Fuerst, C., Yelon, W.: Neutron powder diffraction study of Tb2Fe14B. J. Appl.
Phys. 73, 5884–5886 (1993)
420. Drebov, N., Martinez-Limia, A., Kunz, L., Gola, A., Shigematsu, T., Eckl, T., et al.: Ab initio
screening methodology applied to the search for new permanent magnetic materials. New
J. Phys. 15, 125023 (2013)
421. Liebs, M., Hummler, K., Fähnle, M.: Ab-initio calculation of the effective exchange couplings
in rare-earth—transition-metal intermetallics. J. Magn. Magn. Mater. 124, 239–242 (1993)
422. Andreev, A., Bartashevich, M.: Magnetic properties of Th 2 Fe 14 B and its hydride. J. Less-
Common Met. 167, 107–111 (1990)
423. Sinnema, S., Franse, J., Radwanski, R., Buschow, K., de Mooij, D.: Magnetic measurements
on R2Fe14B and R2Co14B compounds in high fields. Le Journal de Physique Colloques. 46,
C6-301–C6-304 (1985)
424. Verhoef, R., Franse, J.J.M., Menovsky, A.A., Radwanski, R.J., Ji, S., Yang, F., et al.:
High-field magnetization measurements on R2FE14B single-crystals. J. Phys. 49, 565–566
(1988)
425. Li, H.S., Gavigan, J.P., Cadogan, J.M., Givord, D., Coey, J.M.D.: A study of exchange and
crystalline electric-field interactions in ND2CO14B – comparison with ND2FE14B. J. Magn.
Magn. Mater. 72, L241–L246 (1988)
426. Sagawa, M., Fujimura, S., Yamamoto, H., Matsuura, Y., Hirosawa, S.: Magnetic properties of
rare-earth-iron-boron permanent magnet materials. J. Appl. Phys. 57, 4094–4096 (1985)
427. Hirosawa, S., Tokuhara, K., Yamamoto, H., Fujimura, S., Sagawa, M., Yamauchi, H.:
Magnetization and magnetic anisotropy of R2Co14B and Nd2 (Fe1− x Cox) 14B measured
on single crystals. J. Appl. Phys. 61, 3571–3573 (1987)
428. Itoh, T., Hikosaka, K., Takahashi, H., Ukai, T., Mori, N.: Anisotropy energies for Y2Fe14B
and Nd2Fe14B. J. Appl. Phys. 61, 3430–3432 (1987)
429. Burlet, P., Coey, J.M.D., Gavigan, J.P., Givord, D., Meyer, C.: A note on exchange and crystal-
field interactions in R2FE14B compounds – YB2FE14B. Solid State Commun. 60, 723–727
(1986)
430. Meyer, C., Gavigan, J.P., Czjzek, G., Bornemann, H.J.: A study of crystal fields in YB2FE14B
bY YB-174 Mossbauer-spectroscopy. Solid State Commun. 69, 83–86 (1989)
431. Luong, N.H., Thuy, N.P., Tai, L.T., Hien, T.D.: Rare earth magnetocrystalline anisotropy in
R2Fe14B (R= Sm, Er, Tm, Yb). Phys. Status Solidi A. 111, 591–595 (1989)
432. Kou, X., Grössinger, R., Müller, H., Buschow, K.: Anomalous 3d anisotropy of R 2 Fe 14 C
and R 2 Fe 14 B compounds. J. Magn. Magn. Mater. 101, 349–351 (1991)
433. Dennis, K.W., Laabs, F.C., Cook, B.A., Harringa, J.L., Russell, A.M., McCallum, R.W.:
Observations of multi-phase microstructures in R-2(Fe1-xCox)(14)B where R = Nd or Dy.
J. Magn. Magn. Mater. 231, L33–L37 (2001)
434. Leroux, D., Vincent, H., Lheritier, P., Fruchart, R.: Crystallographic and magnetic studies of
ND2CO14B and Y2CO14B. J. Phys. 46, 243–247 (1985)
435. Roux, D.L.: D. Le Roux, Ph.D. thesis, University Grenoble France (1986)
436. Buschow, K.H.J., Demooij, D.B., Sinnema, S., Radwanski, R.J., Franse, J.J.M.: Magnetic and
crystallographic properties of ternary rare-earth compounds of the type R2CO14B. J. Magn.
Magn. Mater. 51, 211–217 (1985)
437. Jacobs, T.H., Denissen, C.J.M., Buschow, K.H.J.: Note on the magnetic-properties of
CE2FE14C. J. Less-Common Met.153, L5–L8 (1989)
438. Denissen, C.J.M., Demooij, B.D., Buschow, K.H.J.: Spin reorientation in ND2FE14C.
J. Less-Common Met. 142, 195–202 (1988)
439. Hellwig, C., Girgis, K., Schefer, J., Buschow, K.H.J., Fischer, P.: Crystal and magnetic-
structure of the permanent-magnet materiaL TB2FE14C. J. Less-Common Met. 169, 147–156
(1991)
440. Deboer, F.R., Verhoef, R., Zhang, Z.D., Demooij, D.B., Buschow, K.H.J.: Magnetic-
properties of ND2FE14C and some related pseudoternary compounds. J. Magn. Magn. Mater.
73, 263–266 (1988)
15 Metallic Magnetic Materials 797

441. Deboer, F.R., Huang, Y.K., Zhang, Z.D., Demooij, D.B., Buschow, K.H.J.: Magnetic and
crystallographic properties of ternary rare-earth compounds of the type R2FE14C. J. Magn.
Magn. Mater. 72, 167–173 (1988)
442. Obbade, S., Isnard, O., Miraglia, S., Fruchart, D., L’heritier, P., Lazaro, F., et al.: Hydro-
genation, crystal structure and magnetic ordering of R 2 Fe 14 C (R≡ Sm, Er, Tm).
J. Less-Common Met. 168, 321–328 (1991)
443. Hellwig, C., Girgis, K., Fischer, P., Buschow, K.H.J., Schefer, J.: Crystal and magnetic-
structure of the permanent magnetic material TM2FE14C. J. Alloys Compd. 184, 175–185
(1992)
444. Hellwig, C., Girgis, K., Schefer, J., Buschow, K.H.J., Fischer, P.: Crystal and magnetic-
structure of the permanent-magnet material LU2FE14C. J. Less-Common Met. 163, 361–368
(1990)
445. Fuerst, C.D., Herbst, J.F.: Formation of R2FE14C compounds (R = Y, CE) by rapid
solidification. J. Appl. Phys. 69, 7727–7730 (1991)
446. Coey, J.M.D., Li, H.S., Gavigan, J.P., Cadogan, J.M., Hu, B.P.: In: Mitchell, I.V., Coey,
J.M.D., Harris, I.R., Hanitsch, R. (eds.) Concerted European Action on Magnets, p. 76.
Elsevier Applied Science Publishing, London (1989)
447. Wolfers, P., Bacmann, M., Fruchart, D.: Single crystal neutron diffraction investigations of the
crystal and magnetic structures of R2Fe14B (R=Y, Nd, Ho, Er). J. Alloys Compd. 317–318,
39–43 (2001)
448. Kido, G., Kato, H., Yamada, M., Nakagawa, Y., Hirosawa, S., Sagawa, M.: Magnetization
anomaly of Sm 2 Fe 14 B single crystal in high magnetic fields. J. Phys. Soc. Jpn. 56, 4635–
4636 (1987)
449. Buschow, K.: New developments in hard magnetic materials. Rep. Prog. Phys. 54, 1123
(1991)
450. Bogé, M., Coey, J., Czjzek, G., Givord, D., Jeandey, C., Li, H., et al.: 3d-4f magnetic
interactions and crystalline electric field in the R 2 Fe 14 B compounds: magnetization
measurements and Mössbauer study of Gd 2 Fe 14 B. Solid State Commun. 55, 295–298
(1985)
451. Loewenhaupt, M., Sosnowska, I.: Exchange and crystal fields in R2Fe14B studied by inelastic
neutron scattering. J. Appl. Phys. 70, 5967–5971 (1991)
452. Ito, M., Yano, M., Dempsey, N.M., Givord, D.: Calculations of the magnetic properties of R
2 M 14 B intermetallic compounds (R= rare earth, M= Fe, Co). J. Magn. Magn. Mater. 400,
379–383 (2016)
453. Allemand, J., Letant, A., Moreau, J., Nozieres, J., De La Bathie, R.P.: A new phase in Nd 2
Fe 14 B magnets. Crystal structure and magnetic properties of Nd 6 Fe 13 Si. J. Less-Comm.
Met. 166, 73–79 (1990)
454. Weitzer, F., Leithe-Jasper, A., Rogl, P., Hiebl, K., Rainbacher, A.: G. experience frustration
with regard to the coupling with the Wiesinger, J. Friedl, FE Wagner. J. Appl. Phys. 75, 7745
(1995)
455. Schrey, P., Velicescu, M.: Influence of Sn additions on the magnetic and microstructural
properties of Nd-Dy-Fe-B magnets. J. Magn. Magn. Mater. 101, 417–418 (1991)
456. Kramer, M., O’connor, A., Dennis, K., McCallum, R., Lewis, L., Tung, L., et al.: Origins of
coercivity in the amorphous alloy Nd-Fe-Al. IEEE Trans. Magn. 37, 2497–2499 (2001)
457. Weitzer, F., Leithe-Jasper, A., Rogl, P., Hiebl, K., Noël, H., Wiesinger, G., et al.: Magnetism
of (Fe, Co)-based alloys with the La 6 Co 11 Ga 3-type. J. Solid State Chem. 104, 368–376
(1993)
458. Schobinger-Papamantellos, P., Buschow, K., Ritter, C.: Magnetic ordering of the Nd 6 Fe 13−
x Ga 1+ x (x= 0, 1) and Pr 6 Fe 13− x Ga 1+ x (x= 0, 1) compounds: a neutron diffraction
study. J. Alloys Compd. 359, 10–21 (2003)
459. Nagata, Y., Kamonji, M., Kurihara, M., Yashiro, S., Samata, H., Abe, S.: Magnetism and
transport properties of Nd 6 Fe 13− x Al 1+ x crystals. J. Alloys Compd. 296, 209–218
(2000)
798 J. P. Liu et al.

460. Kennedy, S., Wu, E., Wang, F., Zhang, P., Yan, Q.: Neutron diffraction study of the magnetic
structure of Pr 6 Fe 11 Al 3. Phys. B. 276, 622–623 (2000)
461. De Groot, C., Buschow, K., De Boer, F.: Magnetic properties of R 6 Fe 13− x M 1+ x
compounds and their hydrides. Phys. Rev. B. 57, 11472 (1998)
462. Schobinger-Papamantellos, P., Buschow, K., De Groot, C., De Boer, F., Ritter, C., Fauth, F.,
et al.: On the magnetic ordering of R 6 Fe 13 X compounds. J. Alloys Compd. 280, 44–55
(1998)
463. Isnard, O., Long, G.J., Hautot, D., Buschow, K., Grandjean, F.: A neutron diffraction and
Mössbauer spectral study of the magnetic spin reorientation in Nd6Fe13Si. J. Phys.: Condens.
Matter. 14, 12391 (2002)
464. Schobinger-Papamantellos, P., Buschow, K., De Groot, C., De Boer, F., Ritter, C.: Magnetic
ordering of the R 6 Fe 13 Sn (R= Nd, Pr) compounds studied by neutron diffraction. J. Magn.
Magn. Mater. 218, 31–41 (2000)
465. De Groot, C., De Boer, F., Buschow, K., Hautot, D., Long, G.J., Grandjean, F.: Magnetic
and Mössbauer spectral properties of the compound Nd 6 Fe 13 Au. J. Alloys Compd. 233,
161–164 (1996)
466. Hu, B.-p., Coey, J., Klesnar, H., Rogl, P.: Crystal structure, magnetism and 57Fe Mössbauer
spectra of ternary RE6Fe11Al3 and RE6Fe13Ge compounds. J. Magn. Magn. Mater. 117,
225–231 (1992)
467. Ruzitschka, R., Reissner, M., Steiner, W., Rogl, P.: Investigation of magnetic order in RE 6
Fe 13 X (RE= Nd, Pr; X= Pd, Sn, Si). J. Magn. Magn. Mater. 242, 806–808 (2002)
468. Iranmanesh, P., Tajabor, N., Pourarian, F.: Magnetostriction effect of Co substitution in the
Nd 6 Fe 13 Si intermetallic compound. Intermetallics. 42, 180–183 (2013)
469. Iranmanesh, P., Tajabor, N., Roknabadi, M.R., Pourarian, F., Brück, E.: Influence of Co
substitution on magnetic properties and thermal expansion of Nd 6 Fe 13 Si intermetallic
compound. Intermetallics. 19, 682–687 (2011)
470. Schobinger-Papamantellos, P., Buschow, K., De Groot, C., De Boer, F., Böttger, G., Ritter,
C.: Magnetic ordering of Pr6Fe13Si and Nd6Fe13Au studied by neutron diffraction. J. Phys.:
Condens. Matter. 11, 4469 (1999)
471. Leithe-Jasper, A., Skomski, R., Qi, Q., Coey, J., Weitzer, F., Rogl, P.: Hydrogen in intermetal-
lic compounds (RE= Pr, Nd; X= Ag, Au, Si, Ge, Sn, Pb). J. Phys.: Condens. Matter. 8, 3453
(1996)
472. Bodak, O., Stepien-Damm, J., Galdecka, E.: Phase equilibria in the ternary systems Pr–Fe–Bi
and Sm–Fe–Bi. J. Alloys Compd. 298, 195–197 (2000)
473. Hautot, D., Long, G.J., Grandjean, F., Buschow, K.: A comparative Mössbauer spectral study
of the electronic and magnetic properties of Nd 6 Fe 13 Ag and Nd 6 Fe 13 AgH 13. J. Alloys
Compd. 388, 159–167 (2005)
474. Wang, F., Zhang, P., Shen, B.-g., Yan, Q., Gong, H.: Transport properties of R6Fe11Al3
compounds (R= La, Nd). J. Appl. Phys. 87, 6043–6045 (2000)
475. Jonen, S., Rechenberg, H.: Magnetoresistance effects at the metamagnetic transition in
R6Fe14− xAlx (R= Nd, La). J. Appl. Phys. 85, 4448–4450 (1999)
476. Coey, J., Qi, Q., Knoch, K., Leithe-Jasper, A., Rogl, P.: Hydrogen induced metamagnetism in
R 6 Fe 13 X compounds. J. Magn. Magn. Mater. 129, 87–97 (1994)
477. Yartys, V., Denys, R., Bulyk, I., Delaplane, R., Hauback, B.: Powder neutron diffraction study
of Nd 6 Fe 13 GaD 12.3 with a filled Nd 6 Fe 13 Si-type structure. J. Alloys Compd. 312,
158–164 (2000)
478. Yartys, V.A., de Boer, F.R., Buschow, K.H.J., Ouladdiaf, B., Brinks, H.W., Hauback, B.C.:
Crystallographic and magnetic structure of Pr6Fe13AuD13. J. Alloys Compd. 356, 142–146
(2003)
479. Schobinger-Papamantellos, P., Ritter, C., Buschow, K.: On the magnetic ordering of Nd 6
Fe 13− x Al 1+ x (x= 1–3) and La 6 Fe 11 Al 3 compounds. J. Magn. Magn. Mater. 260,
156–172 (2003)
480. Leithe-Jasper, A., Rogl, P., Wiesinger, G., Rainbacher, A., Hatzl, R., Forsthuber, M.: A Sn
119 Mössbauer study of RE 6 M 13 Sn (RE= La, Pr, Nd; M= Fe, Co) and their hydrides. J.
Magn. Magn. Mater. 170, 189–200 (1997)
15 Metallic Magnetic Materials 799

481. Nesbitt, E., Wernick, J., Corenzwit, E.: Magnetic moments of alloys and compounds of iron
and cobalt with rare earth metal additions. J. Appl. Phys. 30, 365–367 (1959)
482. Hubbard, W.M., Adams, E., Gilfrich, J.: Magnetic moments of alloys of gadolinium with
some of the transition elements. J. Appl. Phys. 31, S368–S369 (1960)
483. Hoffer, G., Strnat, K.: Magnetocrystalline anisotropy of YCo 5 and Y 2 Co 17. IEEE Trans.
Magn. 2, 487–489 (1966)
484. Strnat, K.J., Hoffer, G.: Air Force Materials Lab. Technical Report AFML-TR-65-446,
Dayton (1965)
485. Strnat, K., Hoffer, G., Olson, J., Ostertag, W., Becker, J.: A family of new cobalt-base
permanent magnet materials. J. Appl. Phys. 38, 1001–1002 (1967)
486. Das, D.K.: Twenty million energy product samarium-cobalt magnet. IEEE Trans. Magn. 5,
214–216 (1969)
487. Benz, M., Martin, D.: Cobalt-samarium permanent magnets prepared by liquid phase
sintering. Appl. Phys. Lett. 17, 176–177 (1970)
488. Ray, A., Strnat, K.: Research and development of rare earth-transition metal alloys as
permanent-magnet materials. Semiannual interim technical report, 1 January–30 June 1972.
Dayton University, Ohio. Research Institute (1971–1973)
489. Ray, A.E., Strnat, K.J.: In: Savitsky, E.I. (ed.) 7th Rare Earth Metals Conference, p. 75. A.A.
Baikov Institute of Meals, Moscow (1972)
490. Mildrum, H., Hartings, M., Strnat, K., Tront, J.: Magnetic properties of the intermetallic
phases Sm $ sub 2$(Co, Fe) $ sub 17$. In: AIP Conference Proceedings No. 10, pp. 618–
622 (1973)
491. Senno, H., Tawara, Y.: Permanent-magnet properties of Sm-Ce-Co-Fe-Cu alloys with com-
positions between 1-5 and 2-17. IEEE Trans. Magn. 10, 313–317 (1974)
492. Ojima, T., Tomizawa, S., Yoneyama, T., Hori, T.: Magnetic properties of a new type of
rare-earth cobalt magnets Sm 2 (Co, Cu, Fe, M) 17. IEEE Trans. Magn. 13, 1317–1319
(1977)
493. Mishra, R.K., Thomas, G., Yoneyama, T., Fukuno, A., Ojima, T.: Microstructure and
properties of step aged rare earth alloy magnets. J. Appl. Phys. 52, 2517–2519 (1981)
494. Massalski, T.B.: In: Massalski, T.B. (ed.) Binary Alloy Phase Diagrams, vol. 2, 2nd edn, p.
1241. ASM International, Materials Park (1990)
495. Khan, Y.: Proceedings of the 11th Rare Earth Research Conference, vol. II, pp. 652–661
(1974)
496. Buschow, K., Van der Goot, A.: Intermetallic compounds in the system samarium-cobalt. J.
Less Common Met. 14, 323–328 (1968)
497. Strnat, J.: Rare earth – cobalt permanent magnets. In: Ferromagnetic Materials – A Handbook
on the Properties of Magnetically Ordered Substances, p. 154, 186, 195. North-Holland
Physics Publishing (1988)
498. Campbell, P.: Permanent Magnet Materials and Their Application, p. 51. Cambridge Univer-
sity Press, Cambridge (1996)
499. Liu, S.: 18th International Workshop on High Performance Magnets, p. 691 (2004)
500. Wang, Y., Li, Y., Rong, C., Liu, J.P.: Sm–Co hard magnetic nanoparticles prepared by
surfactant-assisted ball milling. Nanotechnology. 18, 465701 (2007)
501. Cui, B., Li, W., Hadjipanayis, G.: Formation of SmCo 5 single-crystal submicron flakes and
textured polycrystalline nanoflakes. Acta Mater. 59, 563–571 (2011)
502. Poudyal, N., Liu, J.P.: Advances in nanostructured permanent magnets research. J. Phys. D:
Appl. Phys. 46, 043001 (2012)
503. Hu, D., Yue, M., Zuo, J., Pan, R., Zhang, D., Liu, W., et al.: Structure and magnetic properties
of bulk anisotropic SmCo 5/α-Fe nanocomposite permanent magnets prepared via a bottom
up approach. J. Alloys Compd. 538, 173–176 (2012)
504. Yue, M., Zuo, J., Liu, W., Lv, W., Zhang, D., Zhang, J., et al.: Magnetic anisotropy in bulk
nanocrystalline SmCo5 permanent magnet prepared by hot deformation. J. Appl. Phys. 109,
07A711 (2011)
505. Shen, Y., Leontsev, S.O., Turgut, Z., Lucas, M.S., Sheets, A.O., Horwath, J.C.: Effect of
soft phase on magnetic properties of bulk Sm – Co/<formula formulatype=“inline”> <img
800 J. P. Liu et al.

src=“/images/tex/451.gif” alt=“\alpha ”> </formula> – Fe nanocomposite magnets. IEEE


Trans. Magn. 49, 3244–3247 (2013)
506. Croat, J.J., Herbst, J.F., Lee, R.W., Pinkerton, F.E.: High-energy product Nd-Fe-B permanent
magnets. Appl. Phys. Lett. 44, 148–149 (1984)
507. Croat, J.J., Herbst, J.F., Lee, R.W., Pinkerton, F.E.: Pr-Fe and Nd-Fe-based materials: A new
class of high-performance permanent magnets (invited). J. Appl. Phys.55, 2078–2082 (1984)
508. Sagawa, M., Fujimura, S., Togawa, N., Yamamoto, H., Matsuura, Y.: New material for
permanent magnets on a base of Nd and Fe. J. Appl. Phys. 55, 2083–2087 (1984)
509. Herbst, J.F., Croat, J.J., Pinkerton, F.E., Yelon, W.: Relationships between crystal structure
and magnetic properties in Nd 2 Fe 14 B. Phys. Rev. B. 29, 4176 (1984)
510. Nd-Fe-B Phase Diagram, ASM Alloy Phase Diagrams Database, P. Villars, editor-in-
chief; H. Okamoto and K. Cenzual, section editors; http://www.asminternational.org, ASM
International, Materials Park
511. Herbst, J.F.: ${\mathrm{R}}_{2}$${\mathrm{Fe}}_{14}$B materials: intrinsic properties and
technological aspects. Rev. Mod. Phys. 63, 819–898 (1991)
512. Gutfleisch, O.: Controlling the properties of high energy density permanent magnetic
materials by different processing routes. J. Phys. D: Appl. Phys. 33, R157 (2000)
513. Buschow, K.: Ferromagnetic Materials, vol. 3, 4, p. 1. North-Holland, Amsterdam (1988)
514. Sagawa, M., Hirosawa, S., Yamamoto, H., Fujimura, S., Matsuura, Y.: Nd–Fe–B permanent
magnet materials. Jpn. J. Appl. Phys. 26, 785 (1987)
515. Mishra, R.K.: Microstructure of melt-spun Nd-Fe-B magnequench magnets. J. Magn. Magn.
Mater. 54, 450–456 (1986)
516. Coey, J., Skomski, R.: New magnets from interstitial intermetallics. Phys. Scr. 1993, 315
(1993)
517. GschneidnerJr, K.A., Pecharsky, V., Tsokol, A.: Recent developments in magnetocaloric
materials. Rep. Prog. Phys. 68, 1479 (2005)
518. Moya, X., Kar-Narayan, S., Mathur, N.: Caloric materials near ferroic phase transitions. Nat.
Mater. 13, 439–450 (2014)
519. Moya, X., Defay, E., Heine, V., Mathur, N.D.: Too cool to work. Nat. Phys. 11, 202–205
(2015)
520. Nikitin, S., Myalikgulyev, G., Annaorazov, M., Tyurin, A., Myndyev, R., Akopyan, S.: Giant
elastocaloric effect in FeRh alloy. Phys. Lett. A. 171, 234–236 (1992)
521. Stern-Taulats, E., Planes, A., Lloveras, P., Barrio, M., Tamarit, J.-L., Pramanick, S., et al.:
Barocaloric and magnetocaloric effects in Fe 49 Rh 51. Phys. Rev. B. 89, 214105 (2014)
522. Stern-Taulats, E., Gràcia-Condal, A., Planes, A., Lloveras, P., Barrio, M., Tamarit, J.-L., et al.:
Reversible adiabatic temperature changes at the magnetocaloric and barocaloric effects in
Fe49Rh51. Appl. Phys. Lett. 107, 152409 (2015)
523. Pecharsky, V.K., Gschneidner Jr., K.A.: Giant magnetocaloric effect in Gd 5 (Si 2 Ge 2). Phys.
Rev. Lett. 78, 4494 (1997)
524. Pecharsky, V., Gschneidner, K.: Phase relationships and crystallography in the pseudobinary
system Gd 5 Si 4-Gd 5 Ge 4. J. Alloys Compd. 260, 98–106 (1997)
525. Morellon, L., Blasco, J., Algarabel, P., Ibarra, M.: Nature of the first-order antiferromagnetic-
ferromagnetic transition in the Ge-rich magnetocaloric compounds Gd 5 (Si x Ge 1− x) 4.
Phys. Rev. B. 62, 1022 (2000)
526. Pecharsky, V.K., Gschneidner Jr., K.A.: Tunable magnetic regenerator alloys with a giant
magnetocaloric effect for magnetic refrigeration from {bold {approximately}} 20 to {bold
{approximately}} 290K. Appl. Phys. Lett. 70, 3299 (1997)
527. Yuce, S., Barrio, M., Emre, B., Stern-Taulats, E., Planes, A., Tamarit, J.-L., et al.: Barocaloric
effect in the magnetocaloric prototype Gd5Si2Ge2. Appl. Phys. Lett. 101, 1906 (2012)
528. Hu, F.-X., Shen, B.-G., Sun, J.-R., Cheng, Z.-H., Rao, G.-H., Zhang, X.-X.: Influence of
negative lattice expansion and metamagnetic transition on magnetic entropy change in the
compound LaFe11. 4Si1. 6. Appl. Phys. Lett. 78, 3675–3677 (2001)
529. Shen, B., Sun, J., Hu, F., Zhang, H., Cheng, Z.: Recent progress in exploring magnetocaloric
materials. Adv. Mater. 21, 4545–4564 (2009)
15 Metallic Magnetic Materials 801

530. Hu, F.X.: Magnetic properties and magnetic entropy change of Fe-based La(Fe,M)13
compounds and Ni-Mn-Ga alloys, Ph.D. thesis, (2002)
531. Hu, F., Shen, B., Sun, J., Wang, G., Cheng, Z.: Very large magnetic entropy change near room
temperature. Appl. Phys. Lett. 80, 826–828 (2002)
532. Wang, F., Chen, Y., Wang, G., Shen, B.: Magnetic entropy in La (Fe, Si) 13 compounds. J.
Phys. D: Appl. Phys. 36, 1–5 (2003)
533. Chen, Y.-f., Wang, F., Shen, B.-g., Hu, F.-x., Sun, J.-r., Wang, G.-j., et al.: Magnetic properties
and magnetic entropy change of LaFe11. 5Si1. 5Hy interstitial compounds. J. Phys.: Condens.
Matter. 15, L161 (2003)
534. Fujita, A., Fujieda, S., Hasegawa, Y., Fukamichi, K.: Itinerant-electron metamagnetic transi-
tion and large magnetocaloric effects in La (Fe x Si 1− x) 13 compounds and their hydrides.
Phys. Rev. B. 67, 104416–104427 (2003)
535. Chen, Y.-F., Wang, F., Shen, B.-G., Wang, G.-J., Sun, J.-R.: Magnetism and magnetic entropy
change of LaFe11. 6Si1. 4Cx (x= 0-0.6) interstitial compounds. J. Appl. Phys. 93, 1323–1325
(2003)
536. Shen, J., Gao, B., Zhang, H., Hu, F., Li, Y., Sun, J., et al.: Reduction of hysteresis loss and
large magnetic entropy change in the NaZn13-type LaPrFeSiC interstitial compounds. Appl.
Phys. Lett. 91, 142504–142504 (2007)
537. Manosa, L., González-Alonso, D., Planes, A., Barrio, M., Tamarit, J.-L., Titov, I.S., et al.:
Inverse barocaloric effect in the giant magnetocaloric La–Fe–Si–Co compound. Nat. Com-
mun. 2, 595 (2011)
538. Liu, D., Huang, Q., Yue, M., Lynn, J., Liu, L., Chen, Y., et al.: Temperature, magnetic
field, and pressure dependence of the crystal and magnetic structures of the magnetocaloric
compound Mn 1.1 Fe 0.9 (P 0.8 Ge 0.2). Phys. Rev. B. 80, 174415 (2009)
539. Tegus, O., Brück, E., Buschow, K., De Boer, F.: Transition-metal-based magnetic refrigerants
for room-temperature applications. Nature. 415, 150–152 (2002)
540. Brück, E., Kamarad, J., Sechovsky, V., Arnold, Z., Tegus, O., de Boer, F.: Pressure effects on
the magnetocaloric properties of MnFeP 1− xAsx. J. Magn. Magn. Mater. 310, e1008–e1009
(2007)
541. Thanh, D.C., Brück, E., Tegus, O., Klaasse, J., Gortenmulder, T., Buschow, K.: Magne-
tocaloric effect in MnFe (P, Si, Ge) compounds. J. Appl. Phys. 99, 08Q107 (2006)
542. Hu, F.-X., Shen, B.-g., Sun, J.-R.: Magnetic entropy change in Ni51. 5Mn22. 7Ga25. 8 alloy.
Appl. Phys. Lett. 76, 3460 (2000)
543. Krenke, T., Duman, E., Acet, M., Wassermann, E.F., Moya, X., Mañosa, L., et al.: Inverse
magnetocaloric effect in ferromagnetic Ni–Mn–Sn alloys. Nat. Mater. 4, 450–454 (2005)
544. Liu, J., Gottschall, T., Skokov, K.P., Moore, J.D., Gutfleisch, O.: Giant magnetocaloric effect
driven by structural transitions. Nat. Mater. 11, 620–626 (2012)
545. Mañosa, L., González-Alonso, D., Planes, A., Bonnot, E., Barrio, M., Tamarit, J.-L., et al.:
Giant solid-state barocaloric effect in the Ni-Mn-In magnetic shape-memory alloy. Nat.
Mater. 9, 478–481 (2010)
546. Lu, B., Xiao, F., Yan, A., Liu, J.: Elastocaloric effect in a textured polycrystalline Ni-Mn-In-
Co metamagnetic shape memory alloy. Appl. Phys. Lett. 105, 161905 (2014)
547. Zhao, Y.-Y., Hu, F.-X., Bao, L.-F., Wang, J., Wu, H., Huang, Q.-Z., et al.: Giant negative ther-
mal expansion in bonded MnCoGe-based compounds with Ni2In-type hexagonal structure. J.
Am. Chem. Soc. 137, 1746–1749 (2015)
548. Wang, J.-T., Wang, D.-S., Chen, C., Nashima, O., Kanomata, T., Mizuseki, H., et al.: Vacancy
induced structural and magnetic transition in MnCo1-xGe. Appl. Phys. Lett. 89, 2504 (2006)
549. Trung, N., Biharie, V., Zhang, L., Caron, L., Buschow, K., Brück, E.: From single-to double-
first-order magnetic phase transition in magnetocaloric Mn1− xCrxCoGe compounds. Appl.
Phys. Lett. 96, 162507 (2010)
550. Gercsi, Z., Hono, K., Sandeman, K.G.: Designed metamagnetism in CoMnGe 1− x P x. Phys.
Rev. B. 83, 174403 (2011)
551. Wu, R.-R., Bao, L.-F., Hu, F.-X., Wu, H., Huang, Q.-Z., Wang, J., et al.: Sci. Rep. 5, 18027
(2015)
802 J. P. Liu et al.

552. Kuhrt, C., Schittny, T., Barner, K.: Magnetic b-t phase-diagram of anion substituted mnas –
magnetocaloric experiments. Phys. Status Solidi A. 91, 105–113 (1985)
553. Wada, H., Tanabe, Y.: Giant magnetocaloric effect of MnAs1-xSbx. Appl. Phys. Lett. 79,
3302–3304 (2001)
554. Wada, H., Matsuo, S., Mitsuda, A.: Pressure dependence of magnetic entropy change and
magnetic transition in MnAs1-xSbx. Phys. Rev. B. 79 (2009)
555. Mosca, D.H., Vidal, F., Etgens, V.H.: Strain engineering of the magnetocaloric effect in MnAs
epilayers. Phys. Rev. Lett. 101 (2008)
556. Stewart, G.: Heavy-fermion systems. Rev. Mod. Phys. 56, 755 (1984)
557. Flouquet, J.: On the heavy fermion road. Prog. Low Temp. Phys. 15, 139 (2005)
558. Coleman, P., Maple, B., Millis, A.: Special issue containing papers presented at the Institute
for Theoretical Physics Conference on Non-Fermi Liquid Behaviour in Metals – Santa
Barbara, USA, 17–21 June 1996 – Preface. J. Phys. Conden. Matter. 8, U3–U6 (1996)
559. Stewart, G.: Non-Fermi-liquid behavior in d-and f-electron metals. Rev. Mod. Phys. 73, 797
(2001)
560. Grewe, N., Steglich, F.: In: Gschneidner, K.A., Eyring, L. (eds.) Handbook on thePhysics
and Chemistry of Rare Earths, vol. 14, pp. 343–474. Elsevier Science Publishers, Amsterdam
(1991)
561. Ott, H.R., Fisk, Z.: In: Freeman, A.J., Lander, G.H. (eds.) Handbook on thePhysics and
Chemistry ofActinides, vol. 5, pp. 85–225. Elsevier Science Publishers, Amsterdam (1987)
562. Doniach, S.: The Kondo lattice and weak antiferromagnetism. Phys. B+ C. 91, 231–234
(1977)
563. von Loehneysen, H., Rosch, A., Vojta, M., Woelfle, P.: Fermi-liquid instabilities at magnetic
quantum phase transitions. Rev. Mod. Phys. 79, 1015–1075 (2007)
564. Coleman, P., Pepin, C., Si, Q.M., Ramazashvili, R.: How do Fermi liquids get heavy and die?
J Phys. Conden. Matter. 13, R723–R738 (2001)
565. Si, Q.: Quantum criticality and global phase diagram of magnetic heavy fermions. Phys.
Status Solidi B. 247, 476–484 (2010)
566. Sigrist, M., Ueda, K.: Phenomenological theory of unconventional superconductivity. Rev.
Mod. Phys. 63, 239–311 (1991)
567. Pfleiderer, C.: Superconducting phases of f-electron compounds. Rev. Mod. Phys. 81, 1551–
1624 (2009)
568. White, B.D., Thompson, J.D., Maple, M.B.: Unconventional superconductivity in heavy-
fermion compounds. Phys. C. 514, 246–278 (2015)
569. Andres, K., Graebner, J.E., Ott, H.R.: 4F-virtual-bound-state formation in ceal-3 at low-
temperatures. Phys. Rev. Lett. 35, 1779–1782 (1975)
570. Mattens, W.C.M., Elenbaas, R.A., Boer, F.R.D.: Mixed-valence behavior in intermetallic
compound ybcual. Commun. Phys. 2, 147–150 (1977)
571. Stewart, G.R., Fisk, Z., Wire, M.S.: New Ce heavy-fermion system – CeCu6. Phys. Rev. B.
30, 482–484 (1984)
572. Fujita, T., Satoh, K., Onuki, Y., Komatsubara, T.: Specific-heat of the dense kondo compound
CeCu6. J. Magn. Magn. Mater. 47–8, 66–68 (1985)
573. Besnus, M.J., Kappler, J.P., Lehmann, P., Meyer, A.: Low-temperature heat-capacity, magne-
tization, resistivity of CeRu2SI2, with Y or La substitution. Solid State Commun.55, 779–782
(1985)
574. Haen, P., Flouquet, J., Lapierre, F., Lejay, P., Remenyi, G.: Metamagnetic-like transition in
CeRu2SI2. J. Low Temp. Phys. 67, 391–419 (1987)
575. Lee, W., Shelton, R.: CePtSi: a new heavy-fermion compound. Phys. Rev. B. 35, 5369 (1987)
576. Havela, L., Sechovský, V., Svoboda, P., Diviš, M., Nakotte, H., Prokeš, K., et al.: Heavy
fermion behavior of U2T2X compounds. J. Appl. Phys. 76, 6214–6216 (1994)
577. Yatskar, A., Beyermann, W., Movshovich, R., Canfield, P.: Possible correlated-electron
behavior from quadrupolar fluctuations in PrInA g 2. Phys. Rev. Lett. 77, 3637 (1996)
578. Torikachvili, M., Jia, S., Mun, E., Hannahs, S., Black, R., Neils, W., et al.: Six closely related
YbT2Zn20 (T= Fe, Co, Ru, Rh, Os, Ir) heavy fermion compounds with large local moment
degeneracy. Proc. Natl. Acad. Sci. U.S.A. 104, 9960–9963 (2007)
15 Metallic Magnetic Materials 803

579. Ott, H., Rudigier, H., Delsing, P., Fisk, Z.: Magnetic ground state of a heavy-electron system:
U 2 Zn 17. Phys. Rev. Lett. 52, 1551 (1984)
580. Fisk, Z., Stewart, G., Willis, J., Ott, H., Hulliger, F.: Low-temperature properties of the heavy-
fermion system U Cd 11. Phys. Rev. B. 30, 6360 (1984)
581. Stewart, G., Fisk, Z., Smith, J., Willis, J., Wire, M.: New heavy-fermion system, NpBe 13,
with a comparison to UBe 13 and PuBe 13. Phys. Rev. B. 30, 1249 (1984)
582. Ott, H., Rudigier, H., Felder, E., Fisk, Z., Batlogg, B.: Low-temperature state of U Cu 5:
formation of heavy electrons in a magnetically ordered material. Phys. Rev. Lett. 55, 1595
(1985)
583. Lin, C., Teter, J., Crow, J., Mihalisin, T., Brooks, J., Abou-Aly, A., et al.: Observation of
Magnetic-Field-Induced Superconductivity in a Heavy-Fermion Antiferromagnet: Ce Pb 3.
Phys. Rev. Lett. 54, 2541 (1985)
584. Fisk, Z., Canfield, P., Beyermann, W., Thompson, J., Hundley, M., Ott, H., et al.: Massive
electron state in YbBiPt. Phys. Rev. Lett. 67, 3310 (1991)
585. Kitazawa, H., Schank, C., Thies, S., Seidel, B., Geibel, C., Steglich, F.: A new antifer-
romagnetic heavy fermion compound: CePd 2 Al 3. J. Phys. Soc. Jpn. 61, 1461–1464
(1992)
586. Morosan, E., Bud’ko, S.L., Canfield, P.C., Torikachvili, M.S., Lacerda, A.H.: Thermodynamic
and transport properties of RAgGe (R = Tb-Lu) single crystals. J. Magn. Magn. Mater. 277,
298–321 (2004)
587. Vonlohneysen, H., Pietrus, T., Portisch, G., Schlager, H.G., Schroder, A., Sieck, M., et al.:
Non-fermi-liquid behavior in a heavy-fermion alloy at a magnetic instabilitY. Phys. Rev. Lett.
72, 3262–3265 (1994)
588. Raymond, S., Regnault, L.P., Kambe, S., Mignot, J.M., Lejay, P., Flouquet, J.: Magnetic
correlations in Ce0.925La0.075Ru2Si2. J. Low Temp. Phys. 109, 205–224 (1997)
589. Trovarelli, O., Geibel, C., Mederle, S., Langhammer, C., Grosche, F.M., Gegenwart, P., et al.:
YbRh2Si2: pronounced non-Fermi-liquid effects above a low-lying magnetic phase transition.
Phys. Rev. Lett. 85, 626–629 (2000)
590. Custers, J., Gegenwart, P., Wilhelm, H., Neumaier, K., Tokiwa, Y., Trovarelli, O., et al.: The
break-up of heavy electrons at a quantum critical point. Nature. 424, 524–527 (2003)
591. Paschen, S., Mueller, M., Custers, J., Kriegisch, M., Prokofiev, A., Hilscher, G.,
et al.: Quantum critical behaviour in Ce3Pd20Si6? J. Magn. Magn. Mater. 316, 90–92
(2007)
592. Steglich, F., Aarts, J., Bredl, C.D., Lieke, W., Meschede, D., Franz, W., et al.: Supercon-
ductivity in the presence of strong pauli paramagnetism – CeCu2SI2. Phys. Rev. Lett. 43,
1892–1896 (1979)
593. Stockert, O., Faulhaber, E., Zwicknagl, G., Stusser, N., Jeevan, H.S., Deppe, M., et al.: Nature
of the A phase in CeCu2Si2. Phys. Rev. Lett. 92 (2004)
594. Ott, H.R., Rudigier, H., Fisk, Z., Smith, J.L.: UBe13 – an unconventional actinide supercon-
ductor. Phys. Rev. Lett. 50, 1595–1598 (1983)
595. Ott, H.R., Rudigier, H., Felder, E., Fisk, Z., Smith, J.L.: Influence of impurities and magnetic-
fields on the normal and superconducting states of UBe13. Phys. Rev. B. 33, 126–131 (1986)
596. Stewart, G.R., Fisk, Z., Willis, J.O., Smith, J.L.: Possibility of coexistence of bulk supercon-
ductivity and spin fluctuations in UPt3. Phys. Rev. Lett. 52, 679–682 (1984)
597. Aeppli, G., Bucher, E., Broholm, C., Kjems, J.K., Baumann, J., Hufnagl, J.: Magnetic order
and fluctuations in superconducting UPt3. Phys. Rev. Lett. 60, 615–618 (1988)
598. Fisher, R.A., Kim, S., Woodfield, B.F., Phillips, N.E., Taillefer, L., Hasselbach, K., et al.:
Specific-heat of UPt3 – evidence for unconventional superconductivitY. Phys. Rev. Lett. 62,
1411–1414 (1989)
599. Palstra, T., Menovsky, A., Van den Berg, J., Dirkmaat, A., Kes, P., Nieuwenhuys, G., et al.:
Superconducting and magnetic transitions in the heavy-fermion system U Ru 2 Si 2. Phys.
Rev. Lett. 55, 2727 (1985)
600. Petrovic, C., Pagliuso, P.G., Hundley, M.F., Movshovich, R., Sarrao, J.L., Thompson, J.D.,
et al.: Heavy-fermion superconductivity in CeCoIn5 at 2.3 K. J. Phys.: Condens. Matter. 13,
L337–L342 (2001)
804 J. P. Liu et al.

601. Petrovic, C., Movshovich, R., Jaime, M., Pagliuso, P.G., Hundley, M.F., Sarrao, J.L., et al.: A
new heavy-fermion superconductor CeIrIn5: A relative of the cuprates? Europhys. Lett. 53,
354–359 (2001)
602. Bauer, E.D., Frederick, N.A., Ho, P.C., Zapf, V.S., Maple, M.B.: Superconductivity and heavy
fermion behavior in PrOs4Sb12. Phys. Rev. B. 65 (2002)
603. Vollmer, R., Faisst, A., Pfleiderer, C., von Lohneysen, H., Bauer, E.D., Ho, P.C., et al.: Low-
temperature specific heat of the heavy-fermion superconductor PrOs4Sb12. Phys. Rev. Lett.
90 (2003)
604. Chen, G.F., Ohara, S., Hedo, M., Uwatoko, Y., Saito, K., Sorai, M., et al.: Observation of
superconductivity in heavy-fermion compounds of Ce2CoIn8. J. Phys. Soc. Jpn. 71, 2836–
2838 (2002)
605. Nakatsuji, S., Kuga, K., Machida, Y., Tayama, T., Sakakibara, T., Karaki, Y., et al.:
Superconductivity and quantum criticality in the heavy-fermion system beta-YbAlB(4). Nat.
Phys. 4, 603–607 (2008)
606. Kaczorowski, D., Pikul, A.P., Gnida, D., Tran, V.H.: Emergence of a superconducting state
from an antiferromagnetic phase in single crystals of the heavy fermion compound Ce2PdIn8.
Phys. Rev. Lett. 103 (2009)
607. Geibel, C., Thies, S., Kaczorowski, D., Mehner, A., Grauel, A., Seidel, B., et al.: A new
heavy-fermion superconduCTOR – UNI2AL3. Z. Phys. B: Condens. Matter. 83, 305–306
(1991)
608. Geibel, C., Schank, C., Thies, S., Kitazawa, H., Bredl, C.D., Bohm, A., et al.: Heavy-fermion
superconductivity at Tc=2K in the antiferromagnet UPd2AL3. Z. Phys. B: Condens. Matter.
84, 1–2 (1991)
609. Bauer, E., Hilscher, G., Michor, H., Paul, C., Scheidt, E.W., Gribanov, A., et al.: Heavy
fermion superconductivity and magnetic order in noncentrosymmetric CePt3Si. Phys. Rev.
Lett. 92 (2004)
610. Prokleska, J., Kratochvilova, M., Uhlirova, K., Sechovsky, V., Custers, J.: Magnetism,
superconductivity, and quantum criticality in the multisite cerium heavy-fermion compound
Ce3PtIn11. Phys. Rev. B. 92 (2015)
611. Jaccard, D., Behnia, K., Sierro, J.: Pressure-induced heavy fermion superconductivity of
CeCu2Ge2. Phys. Lett. A. 163, 475–480 (1992)
612. Fisher, R.A., Emerson, J.P., Caspary, R., Phillips, N.E., Steglich, F.: The low-temperature
specific-heat of CeCu2Ge2 at 0-kbar and 9.5-kbar. Phys. B. 194, 459–460 (1994)
613. Grosche, F.M., Julian, S.R., Mathur, N.D., Lonzarich, G.G.: Magnetic and superconducting
phases of CePd2Si2. Phys. B. 223–24, 50–52 (1996)
614. Steeman, R.A., Frikkee, E., Helmholdt, R.B., Menovsky, A.A., Vandenberg, J., Nieuwenhuys,
G.J., et al.: CePd2SI2 – a reduced-moment antiferromagnet. Solid State Commun. 66, 103–
107 (1988)
615. Movshovich, R., Graf, T., Mandrus, D., Thompson, J.D., Smith, J.L., Fisk, Z.: Superconduc-
tivity in heavy-fermion CeRh2Si2. Phys. Rev. B. 53, 8241–8244 (1996)
616. Walker, I.R., Grosche, F.M., Freye, D.M., Lonzarich, G.G.: The normal and superconducting
states of CeIn3 near the border of antiferromagnetic order. Phys. C. 282, 303–306 (1997)
617. Hegger, H., Petrovic, C., Moshopoulou, E.G., Hundley, M.F., Sarrao, J.L., Fisk, Z., et al.:
Pressure-induced superconductivity in quasi-2D CeRhIn5. Phys. Rev. Lett. 84, 4986–4989
(2000)
618. Kimura, N., Ito, K., Saitoh, K., Umeda, Y., Aoki, H., Terashima, T.: Pressure-induced
superconductivity in noncentrosymmetric heavy-fermion CeRhSi3. Phys. Rev. Lett. 95 (2005)
619. Sugitani, I., Okuda, Y., Shishido, H., Yamada, T., Thamhavel, A., Yamamoto, E., et al.:
Pressure-induced heavy-fermion superconductivity in antiferromagnet CeIrSi3 without inver-
sion symmetry. J. Phys. Soc. Jpn. 75 (2006)
620. Saxena, S.S., Agarwal, P., Ahilan, K., Grosche, F.M., Haselwimmer, R.K.W., Steiner, M.J.,
et al.: Superconductivity on the border of itinerant-electron ferromagnetism in UGe2. Nature.
406, 587–592 (2000)
15 Metallic Magnetic Materials 805

621. Aoki, D., Huxley, A., Ressouche, E., Braithwaite, D., Flouquet, J., Brison, J.P., et al.:
Coexistence of superconductivity and ferromagnetism in URhGe. Nature. 413, 613–616
(2001)
622. Kobayashi, T.C., Fukushima, S., Hidaka, H., Kotegawa, H., Akazawa, T., Yamamoto, E.,
et al.: Pressure-induced superconductivity in ferromagnet UIr without inversion symmetry.
Phys. B. 378–80, 355–358 (2006)
623. Huy, N.T., Gasparini, A., de Nijs, D.E., Huang, Y., Klaasse, J.C.P., Gortenmulder, T., et al.:
Superconductivity on the border of weak itinerant ferromagnetism in UCoGe. Phys. Rev. Lett.
99 (2007)

J. Ping Liu received his PhD in Physics from the University


of Amsterdam, the Netherlands. He is currently a Distinguished
Professor at the University of Texas at Arlington, United States
and has worked in several institutions in China, Europe and the
United States in research and development of permanent magnets,
magnetic nanoparticles and nanocomposite materials.

Matthew Willard received his PhD from Carnegie Mellon Uni-


versity in 2000. He worked at the U.S. Naval Research Laboratory
in Washington, DC for 12 years prior to joining the Materials
Science and Engineering faculty at Case Western Reserve Univer-
sity in Cleveland, OH. His expertise includes physical metallurgy,
alloy design of soft magnetic alloys, and melt processing of
alloys.

Wei Tang received his PhD from the Northwestern Polytechnical


University of China in 1996. He works now at Division of
Materials Science and Technology, Ames Laboratory of USDOE
as a Research Scientist. His research scope and interest focus on
the studies of magnetic materials and development of permanent
magnets with high performance.
806 J. P. Liu et al.

Ekkes Brück received his PhD from University of Amsterdam


in 1991. He worked at CMR Stanford University, University of
Amsterdam and Federal University of Santa Catharina, Brazil.
Currently he holds the chair of Fundamental Aspects of Mate-
rials and Energy at Delft University of Technology. Employing
microscopic and macroscopic techniques, main research interests
are materials for renewable energy and energy saving.

Enke Liu received his PhD from Institute of Physics (IOP),


Chinese Academy of Sciences (CAS) at Beijing in 2012. He
works at State Key Lab. of Magnetism, IOP, CAS. He now
works on the experimental realization and energy-band design of
magnetic martensitic alloys and magnetic topological semimetals
including the families of hexagonal MM’X, all-d-metal Heuslers
and Shandites.

Claudia Felser studied chemistry and physics at the University of


Cologne and completed her doctorate in 1994. After postdoctoral
fellowships at the Max Planck Institute (MPI) and the CNRS in
Nantes, she joined the University of Mainz and became a Full
Professor in 2003. She has been director and scientific member at
the MPI for Chemical Physics of Solids since December 2011.

Olivier Isnard received his PhD from Univ. Joseph Fourier and
Institut Laue Langevin Grenoble France in 1993. Now Profes-
sor of Physics at Néel Institute, University Grenoble Alpes, he
worked at the University of Missouri in USA, University of
Amsterdam NL. Doctor Honoris Causa of both Babes Bolyai and
Technical Universities in Cluj Romania, he works on magnetism
of rare-earth transition metal intermetallics.
15 Metallic Magnetic Materials 807

Sam Liu received his Ph.D. from the University of Dayton in


1989. He worked at the University of DaytonMagnetics Lab-
oratory from 1989 to 2011. His research interests in the US
were primarily focused on high-temperature metal-semiconductor
contacts, high-temperature Sm-Co magnets, and nanocrystalline
and nanocomposite rare earth permanent magnet materials.

J. F. Herbst received his PhD in physics from Cornell University


(Ithaca, NY) in 1974. During a career at the General Motors
Research and Development Center he participated in a variety of
new materials projects, including (1) the discovery of Nd2 Fe14 B
and the development of permanent magnets based on that fasci-
nating compound, and (2) hydrogen storage media for fuel cell
vehicles.

Fengxia Hu is a Professor at Institute of Physics, Chinese


Academy of Science since 2008. She received her Ph.D in Physics
at Institute of Physics, Chinese Academy of Science in 2002 with
major of condensed matter physics. Her current interests include
magnetocaloric effect, and magnetic refrigerant applications of
various intermetallics, such as Fe-based NaZn13-type materials,
NiMn-based Heusler alloys, and etc.
808 J. P. Liu et al.

Anne de Visser is associate professor at the University of


Amsterdam. His research is directed towards the understanding
of collective phenomena in novel quantum materials, notably
strongly correlated electron systems and topological materials.
He is a pioneer in the field of heavy-fermion physics, quantum
phase transitions and unconventional superconductivity, with an
expertise in measuring transport, magnetic and thermal properties
under extreme conditions.
Metallic Magnetic Thin Films
16
D. Wu and X.-F. Jin

Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 810
GaAs (001) Substrates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 811
Structure and Magnetism of Magnetic 3d Transition Metals . . . . . . . . . . . . . . . . . . . . . . . . . . 812
Cr/GaAs (001) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 812
Mn/GaAs (001) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 814
Fe/Cu (001) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 816
Co/GaAs (001) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 819
Ni/GaAs (001) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 824
Structure and Magnetism of Magnetic 3d Transition Metal Alloys . . . . . . . . . . . . . . . . . . . . . 826
Py/GaAs (001) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 826
Cox Mn1−x /GaAs (001) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 829
Fex Cu1−x /GaAs (001) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 833
Fex Pd1−x /Cu (100) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 837
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 840

Abstract

Molecular beam epitaxy (MBE) is an advanced technique to grow single-


crystalline films. In this chapter we survey the epitaxial growth of 3d transition
magnetic metal and their alloy films via MBE. Several metastable structure
phases are obtained in the ultrathin film. For example, the body-centered cubic

D. Wu
National Laboratory of Solid State Microstructures and Department of Physics, Nanjing
University, Nanjing, People’s Republic of China
e-mail: dwu@nju.edu.cn
X.-F. Jin ()
Department of Physics and State Key Laboratory of Surface Physics, Fudan University, Shanghai,
People’s Republic of China
e-mail: xfjin@fudan.edu.cn

© Springer Nature Switzerland AG 2021 809


J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_19
810 D. Wu and X.-F. Jin

(bcc) phase of Ni, which does not exist in nature, can be epitaxially grown on
GaAs (001). It is found that the interface structure and film growth temperature
play a crucial role in determining the film structure and magnetism. With these
metastable structure phase and precisely controlled composition of 3d transition
magnetic metal and their alloy films, we explore several interesting physics such
as the origin of the extremely small magnetocrystalline anisotropy in permalloy
and the correlation between the structure and magnetism in Co1−x Mnx alloys.

Introduction

Significant progress has been made in the epitaxial growth of magnetic metallic
films (MMFs) by molecular beam epitaxy (MBE) over the past four decades.
Nowadays it is routinely growing single-crystalline MMFs on different substrates
with submonolayer control. This allows us to understand the fascinating funda-
mental properties of the MMFs, which leads to important applications in current
information technologies. A remarkable example is the discovery of the giant
magnetoresistance (GMR) effect in metallic magnetic multilayers (e.g., Fe/Cr or
Co/Cu superlattices) [1, 2] and the employment of the GMR in the read head of
hard disks. Since then, a variety of spin transport phenomena were discovered and
applied to spintronic devices.
There are four ferromagnetic (FM) elementary metals, Fe, Co, Ni, and Gd. Fe,
Co, and Ni and their alloys have been dominant materials in magnetism in both
fundamental studies and applications, partially because their Curie temperature
(Tc) is much higher than room temperature. Cr and Mn are the only two anti-
ferromagnetic (AFM) metallic elements. According to the Slater-Pauling model
[3, 4], the magnetic state is related to the occupation of d band around the Fermi
level. These five magnetic elementary metals are all 3d transition metals with
atomic numbers continuously changing from 24 to 28. The binary metal alloys
with precisely controlled composition provide a model system to tune the Fermi
level and manipulate the corresponding magnetic properties. Therefore, they and
their alloys are the ideal platforms for a fundamental understanding of magnetism
in the thin film state. Moreover, the magnetic properties of the films such as the
magnetic anisotropy and the magnetic order are substantially different from these of
bulk. A comprehensive and systematic investigation of the 3d MMFs is important.
The recent technological applications of spintronics are mostly based on the 3d
MMFs, and spintronics is also an important driving force for their investigation. For
example, the spin-orbit torque phenomenon which can switch the magnetic moment
by a dc current occurs in AlOx /Co/Pt with Co layers a few monolayers thick [5].
The 3d transition metals and alloys have a variety of crystallographic and
magnetic phases in bulk, some of which only exist at high pressure or high
temperature. In addition to these thermodynamically stable phases, metastable
structures could be epitaxially grown on single-crystalline substrates. In this chapter
we focus on the epitaxial growth and magnetic properties of the magnetic 3d
transition metal and their alloy films.
16 Metallic Magnetic Thin Films 811

GaAs (001) Substrates

Among several film growth methods such as MBE, magnetron sputtering, and
pulsed laser deposition (PLD), MBE is the most powerful technique to grow the
highest-quality MMFs. GaAs (001) substrates are widely used for the epitaxial
growth of 3d transition metals and their alloys. The cubic lattice constant is
0.565 nm. The top view of the unreconstructed GaAs (001) together with the corre-
sponding Miller indices is schematically shown in Fig. 1a. If a cubic-structure film is
grown on GaAs (001) surface, there are two kinds of epitaxial growth geometry: (i)
the face-centered cubic (fcc) structure with (001)[110]film //(001)[100]GaAs , in which
the epitaxial film rotates 45◦ around the surface normal, and (ii) the body-centered
cubic (bcc) structure with (001)[100]film //(001)[100]GaAs , as schematically shown
in Fig. 1b, c, respectively. The lattice constants are 0.399 nm for the fcc structure
and 0.283 nm for the bcc structure without distortion. Lattice misfits are relatively
small for 3d transition metals and alloys, meaning that GaAs is the ideal substrate
for these metals. For example, the lattice misfits are 0.1% for bcc Co/GaAs film
and − 8.6% for fcc Mn/GaAs film.
Te-doped GaAs (001) single crystal wafers can be prepared by two different
procedures after cleaning in acetone, water, and alcohol in an ultrasonic bath in

Fig. 1 Schematic top view of the unreconstructed GaAs (001), together with rotated fcc
structure film with (001)[110]film //(001)[100]GaAs and unrotated bcc structure film with
(001)[100]film //(001)[100]GaAs [6]. LICENSE #: 4812820853366
812 D. Wu and X.-F. Jin

sequence. One is a standard method for most substrates. The GaAs substrate is

bombarded by 800 eV argon ions and then annealed at 500 C for several cycles
in the ultrahigh vacuum (UHV). The other is a chemical etching method, using
H2 SO4 :H2 O2 :H2 O = 5:1:1 to etch the substrate. Then it is loaded into the UHV

system and flashed to 580 C. The observation of a clear (4 × 1) low-energy electron
diffraction (LEED) or reflection high-energy electron diffraction (RHEED) pattern
indicates a clean surface. Auger spectra show the surface is free of carbon and
oxygen contamination. We find that the latter procedure gives a better substrate for
epitaxial growth.

Structure and Magnetism of Magnetic 3d Transition Metals

Cr/GaAs (001)

Chromium has a bcc structure in bulk with a lattice constant of 0.2885 nm. The
lattice mismatch between GaAs (001) substrate and bcc structure Cr is less than
1.9%. The Cr film growth on GaAs (001) is expected to be bcc with the epitaxial
relationship: (001)[100]Cr //(001)[100]GaAs and (001)[100]Cr //(001)[100]GaAs , as
shown in Fig. 1c.
The structure of Cr films grown on the GaAs (001) substrate is dependent on the
growth temperature [7]. Figure 2a presents the typical RHEED pattern for a clean
GaAs (001) surface, with an electron beam along the GaAs [110] direction. In the
initial stage of the film growth at room temperature, the RHEED pattern shows some
polycrystalline circles, seen in Fig. 2b. After the thickness is beyond ∼1.2 nm, a new
diffraction pattern emerges and persists as long as Cr is being deposited. The typical
pattern at this stage is shown in Fig. 2c.
The previous study found that the orientation of the Cr/GaAs (001) film
is (112)[110]Cr //(001)[110]GaAs and (112)[111]Cr //(001)[110]GaAs , instead of the
geometry shown in Fig. 1c [8]. To determine the epitaxial orientation, we schemati-
cally draw these two growth geometries in Fig. 3a, b together with the corresponding
Miller indices. There are two types of structures labeled as (I) and (II) for (112)
plane growth geometry according to the structure symmetry. The RHEED patterns
of type I and type II structures are calculated and shown in Fig. 2f, i, respectively.
Comparing with the experimental results (Fig. 2c), we can immediately determine
that only type I structure is formed. This result is further confirmed by X-ray
diffraction (XRD) [7]. This growth geometry has not been found in other 3d
transition metals and their alloys. It is still not understood why the (112) plane
growth occurs since the lattice mismatch is as high as 25% along [110] direction
or why only type I structure can be grown from the symmetry point of view.
For growth temperatures between 130 and 160◦ C, the RHEED pattern is different
from the room temperature growth case after the initial deposition of ∼1.4 nm, as
shown in Fig. 2e. The growth geometry of Fig. 3a is calculated and shown in Fig. 2h,
in good agreement with the experimental pattern of Fig. 3e. Therefore, the growth
16 Metallic Magnetic Thin Films 813

Fig. 2 RHEED patterns of Cr grown on the GaAs (001) substrate at different growth temperatures
and thickness, taken with the incident electron beam along the substrate [110] direction: (a) clean
GaAs (001); (b) 1.0 nm, 25◦ C; (c) 5.0 nm, 25◦ C; (d) 1.6 nm, 100◦ C; (e) 4.0 nm, 160◦ C; (i–i)
calculated diffraction patterns explained in the text [7]

geometry of Fig. 3a is unambiguously determined to be an unrotated bcc structure.


Again, this result is further confirmed by XRD [7].
For growth temperatures between 70 and 130◦ C, the RHEED pattern is a mixture
of Fig. 2c, e, as shown in Fig. 2d. Figure 2g shows the mixture of the calculated
pattern of Fig. 2f, h. It is clearly matched to the experimental results, indicating that
the two growth geometries coexist in the Cr/GaAs (001) films.
814 D. Wu and X.-F. Jin

Fig. 3 Epitaxial relationships between Cr and GaAs: (a) (001)[100]Cr //(001)[100]GaAs and
(001)[100]Cr //(001)[100]GaAs ; (b) shadow I, (112)[110]Cr //(001)[110]GaAs and (112)[111]Cr //
(001)[110]GaAs ; shadow II, (112)[110]Cr //(001)[110]GaAs and (112)[111]Cr //(001)[110]GaAs [7].
LICENSE #: 4812840298511

Mn/GaAs (001)

Several stable structure phases are available for bulk Mn [9]. The α-phase, which
is a complex cubic structure with 58 atoms per unit cell, is stable below 1000 K.
It is antiferromagnetic (AFM) below 95 K. The β-phase, which is also a complex
cubic structure with 20 atoms per unit cell, is stable between 1000 and 1364 K and
can be retained at room temperature by quenching. The β-phase is nonmagnetic.
The γ -phase, which has the fcc structure, is stable between 1364 and 1410 K. The
δ-phase, which is bcc structure, is stable between 1410 and 1450 K. Both γ - and
δ-phases exist well beyond any magnetic ordering temperature. The metal-stable
phases of Mn were successfully grown on Pd (001) and Ag (001) surfaces with
distorted fcc structure [10, 11]. A nearly fcc structure with 3-monolayer (ML) Mn
layers can be obtained in the (Mn/Ag)n superlattice on Ag (001) [12]. Epitaxial
16 Metallic Magnetic Thin Films 815

growth on GaAs (001) provides another approach to stabilize metastable Mn at


room temperature. The lattice constants of bcc and fcc structure Mn are estimated
to be 0.289 nm and 0.365 nm, respectively, calculated by a simple constant volume
approximation based on the Wigner-Seitz radius of α-phase Mn (0.143 nm) [11].
The lattice mismatches between GaAs (001) and bcc and fcc Mn are about 2%
and 9%.
Figure 4 presents the RHEED patterns of clean GaAs (001) surface and
8-nm-thick Mn/GaAs (001) film grown at 400 K with the incident electron beam
along different directions [13]. The spotlike RHEED patterns reflect the rough
surface and is the projection of the three-dimensional reciprocal lattice. The RHEED
pattern remains the same when the sample is rotated 90◦ C, meaning that the single-
crystalline film is the cubic structure . In comparison with the RHEED pattern of
GaAs (001) (Fig. 2a), the lattice constant of Mn film is determined to be 0.368 nm.
The RHEED pattern of the sample rotated 45◦ C, i.e., electron beam along [100]
direction, is shown in Fig. 4b [13].
The RHEED patterns of the rotated fcc Mn and unrotated bcc Mn are deduced
from the crystal structure in Fig. 1 with electron beam along the [110] and the [100]
directions. The results are shown in Fig. 5 [13]. The deduced RHEED patterns of the
fcc structure is in good agreement with the experimental patterns (Fig. 4), suggesting
that fcc structure Mn films are obtained.
Figure 6 shows the XRD spectra of three Mn/GaAs (001) films grown at 300,
400, and 500 K, respectively [14]. The structure of the film grown at 300 K is
polycrystalline α-phase, as shown in Fig. 6a. A (002) peak of the fcc structure is
clearly present in XRD, as shown in Fig. 6b, confirming the RHEED result. The
measured out-of-plane lattice constant of 0.362 nm is consistent with the value

Fig. 4 RHEED patterns for (a) the clean GaAs (001) surface and (b), (c) fcc Mn thin films of 8 nm
grown on the GaAs (001) substrate with different electron incident directions, respectively [13].
LICENSE #: 4812840822141
816 D. Wu and X.-F. Jin

Fig. 5 RHEED diffraction


patterns calculated for the
rotated fcc Mn (001) and the
unrotated bcc Mn (001) on
the GaAs (001) surface [13].
LICENSE #: 4812840822141

determined by the RHEED patterns. In addition to the fcc structure, we observe


another structure with an out-of-plane lattice constant of 0.207 nm. We attribute
it to the Mn2 As-type structure (003) peak with in-plane lattice constant a and b
of 0.380 nm and out-of-plane lattice constant c of 0.627 nm [15]. This Mn2 As-type
structure is a mixture of Mn-Ga-As and serves as a template for the fcc Mn to reduce
the large lattice mismatch between the fcc Mn and GaAs (001).
To confirm the structures, a cross-sectional high-resolution transmission electron
microscope (HRTEM) is shown in Fig. 6d [14]. Obviously, the fcc Mn is grown on
the interfacial Mn2 As-type phase. The lattice constants of these two structures are
determined by HRTEM: a = b = 0.380 nm, c = 0.363 nm for the fcc-Mn film, and
a = b = 0.380 nm, c = 0.208 nm for the fcc-Mn Mn2 As-type phase, consistent with
the RHEED and XRD results. This finding unambiguously confirms the formation
of the Mn2 As-type phase.
For the film grown at 500 K, two crystallographic phases are observed in XRD
similar to that of 400 K, as shown in Fig. 6c. We can identify these two structures as
“fcc-Mn-type” and “Mn2 As-type” phases. The “fcc-Mn-type” is not pure Mn, but it
is the mixture of Mn-Ga-As. The out-of-plane lattice constant of the “fcc-Mn-type”
phase is 0.375 nm, slightly larger than that of the pure fcc Mn.

Fe/Cu (001)

Iron exhibits a rich variety of magnetic phases. Fe is bcc and FM in bulk. The fcc
phase is stable above 910◦ C. It is predicted to be nonmagnetic, AFM, FM, or order
as a spin density wave (SDW), depending on the lattice constant [16–22]. The fcc-Fe
16 Metallic Magnetic Thin Films 817

Fig. 6 XRD spectra for Mn on GaAs (001), prepared at different growth temperature and film
thickness: (a) 300 K, 20 nm Mn; (b) 400 K, 10 nm Mn; and (c) 500 K, 20 nm Mn. (d) HRTEM
image for 10 nm Mn on GaAs (001) grown at 400 K [14]. LICENSE #: 4812841356913

phase can be stabilized in small particles (d ∼ 50 nm) in a Cu matrix, and it is an


AFM incommensurate SDW [23, 24].
Epitaxial growth of Fe on Cu (001) at 250–300 K can stabilize the bcc and
fcc phases, depending on the film thickness [25–30]. For film thickness below ∼4
monolayers (ML) in region I, it grows in the FM face-centered tetragonal (fct) phase.
For a film thickness between 5 and 11 ML (region II), it is the AFM fcc phase
covered by an FM fct Fe surface layer. However, the energy of SDW is lower than
that of the collinear type 1 AFM state. Its magnetic state is under debate. For the
film thickness above 11 ML, it becomes the FM bcc phase.
A clear RHEED oscillation can be observed for Fe grown on Cu (001) [31],
indicating a layer-by-layer growth mode. It is noticed that quite different RHEED
oscillations are found in the literature [Refs. 28, 30]. We found that the oscillation
mode changes while the electron beam is measured along different azimuthal angles.
Based on this fact, the previously reported oscillation modes can be well reproduced,
as shown in Fig. 7 [31]. The film thickness is further accurately measured by in situ
818 D. Wu and X.-F. Jin

Fig. 7 (a) and (b) Two types of RHEED oscillations for Fe/Cu (001) grown at 300 K [31].
LICENSE #: RNP/20/APR/024963

scanning tunneling microscopy and a quartz crystal monitor. It is very critical to


verify that the studied phase is fcc Fe. A martensitic phase transition from fcc to
bcc occurs at about 10–12 ML, which can be observed by a scanning tunneling
microscope (STM) [32]. Based on this fact, here, we use the observation of the
martensitic phase transition by STM and low-energy electron diffraction (LEED) as
a criterion to judge the realization of fcc Fe.
Figure 8 shows the polar Kerr rotation angle at remanence as a function of film
thickness measured at 70 K. A minimum is clearly seen at ∼7 ML and the signal
does not drop down at 8.5 and 9 ML. These results cannot be explained by a collinear
type 1 antiferromagnet, excluding the type 1 AFM phase of fcc Fe. Figure 9 shows
three representative polar magneto-optic Kerr effect (MOKE) signals at remanence
versus temperature for 6, 7, and 9 ML of Fe on Cu (001), respectively. The total
moment of 7 ML (odd layer) is smaller than that of 6 and 8 ML (even layers) below
∼200 K, which is the effective ordering temperature Te of the AFM underlayers
[30]. The coercivity Hc increases significantly when the temperature drops below
Te, as seen in the inset of Fig. 9, and the hysteresis curves are shown in Fig. 9a–e.
The enhancement of Hc is due to the exchange coupling between the top two FM
surface layers and the bulk AFM underlayers below Te. No exchange bias loops are
observed in these experiments, possibly due to the weak pinning effect of the thin
AFM layer.
We propose a SDW model to explain the data [31]. A SDW with Sz = Sz0
cos[q(z−z0 )] is assumed, where Sz is the z component of the magnetic moment,
Sz0 is the normalization constant, q is the wave vector of SDW, z is the position
in space along the z direction, and z0 is the initial phase term determined by the
choice of the origin. Since the total magnetic moment increases below Te, the
Fe magnetic moments at the FM/AFM interface must be ferromagnetically rather
than antiferromagnetically coupled to each other. It means that z0 = 0 if choosing
the origin at the topmost AFM plane, i.e., the third layer from the top. The inset
16 Metallic Magnetic Thin Films 819

Fig. 8 Polar MOKE as a


function of film thickness
measured at 70 K together
with the fitting curve using
the incommensurate SDW
with Sz = Sz cos(qz) and
q = 2π/2.7d (solid line) [31].
LICENSE #:
RNP/20/APR/024963

Fig. 9 Temperature-
dependent magnetization as a
function of temperature for 6,
7, and 8 ML Fe on Cu (100),
respectively; curves (a–e) are
some representative
hysteresis loops at different
temperatures; (f) coercivity
versus temperature for 8 ML
Fe/Cu (001) [31]. LICENSE
#: RNP/20/APR/024963

of Fig. 10 shows the incommensurate SDW along z direction. The fitting of the
experimental data to an SDW yields q = 2π/2.7d, where d is the interlayer distance
of fcc Fe, as shown by the solid line in Fig. 8. The good fitting strongly suggests that
the magnetic structure is SDW. The magnetic configurations of 6, 7, 8, and 9-ML
Fe on Cu (001) can be constructed one by one as schematically shown in Fig. 10.

Co/GaAs (001)

The hexagonal close-packed (hcp) and fcc structures are, respectively, the stable and
metastable phases of Co. Although the bcc structure is not available in bulk, it was
realized by epitaxial growth on GaAs (110) substrates. A negative cubic magnetic
anisotropy K1 and a strong in-plane uniaxial anisotropy Ku were proposed [33]. The
negative K1 denotes the easy axes along bcc Co <110> directions. However, it was
pointed out that bcc Co phase is not a true metastable phase but a force-induced
phase [34]. The bcc Co phase grown on GaAs (001) substrate was first reported by
820 D. Wu and X.-F. Jin

Fig. 10 Magnetic structures


proposed for 6, 7, 8, and
9-ML Fe on Cu (100); the
inset gives the
layer-dependent magnetic
moments for fcc Fe along the
z direction, z(d) = 0
corresponding to the first
AFM layer. (Note: all the
moments drawn here are
lying in the planes parallel to
the front plane of the structure
section) [31]. LICENSE #:
RNP/20/APR/024963

Blundell et al. [35] They found a positive K1 for 4-nm Co/GaAs (001), contrary to
a previous report. Another group reported that Co films grown on GaAs (001) were
not bcc but two-domain hcp, resulting in positive K1 [36]. Theoretical calculation
predicted that the surface layer contributes to a large positive K1 [37].
The structure of Co grown on GaAs (001) depends on film thickness and growth
temperature. Figure 11a shows a typical RHEED pattern for a clean GaAs (001)
surface, with an electron beam along the [110] direction. A new RHEED pattern
appears at ∼0.2 nm for Co films grown at 150◦ C. This new pattern is clearly seen
after ∼0.8 nm, shown in Fig. 11d [38]. The pattern remains the same by rotating the
electron beam 90◦ to the [110] direction, indicating a fourfold in-plane geometry.
This pattern is the same as the calculated bcc structure RHEED pattern (Fig. 11g).
Moreover, we compare the RHEED pattern of the bcc Fe [39] (Fig. 11b) and fcc Mn
(Fig. 11c) epitaxially grown on GaAs (001). From the similarity between Fig. 11b, d,
again, we can immediately conclude that the structure is bcc.
When the film thickness is beyond ∼2 nm, a new RHEED pattern develops
together with the bcc pattern, shown in Fig. 11e. If we consider the new phase
is hcp Co proposed earlier [36], the calculated pattern of the mixture of the bcc
and hcp structure (Fig. 11h) is consistent with the experimental result (Fig. 11e).
The growth geometry is determined to be (1210)[0001]Co //(001)[110]GaAs and
(1210)[0001]Co //(001[110]GaAs .
When the film thickness exceeds ∼6 nm, the RHEED pattern is dominated by
the hcp structure, as shown in Fig. 11f. Interestingly, the pattern remains the same
after rotating the electron beam 90◦ , contrary to the sixfold symmetry of the hcp
structure. This implies the fourfold symmetry pattern might actually come from the
two perpendicularly oriented hcp Co domains, with [0001]Co //(001)[110]GaAs and
16 Metallic Magnetic Thin Films 821

Fig. 11 RHEED patterns for different thin films grown on the GaAs (001) substrate, taken with
the incident electron beam along the substrate [110] direction: (a) clean GaAs (001), (b) 5.6-nm
bcc Fe, (c) 8.0-nm fcc Mn, (d) 1.5-nm bcc Co, (e) 3.5-nm bcc and hcp Co, (f) 15-nm hcp Co, and
(g–i) calculated diffraction patterns corresponding to (d–f), respectively. ◦ and • correspond to the
two hcp domains described in the text [38]. LICENSE #: RNP/20/APR/024982
822 D. Wu and X.-F. Jin

[0001]Co //(001[110]GaAs together [36]. The calculated pattern is shown in Fig. 11i
according to the two-domain structure, in good agreement with the experimental
result (Fig. 11f). In particular, the intensity of the second and fourth columns from
right to left are much stronger than they ought to be, owing to the overlap of the
two-domain diffraction.
The structure of the Co/GaAs (001) film is directly studied by HRTEM, as shown
in Fig. 12 [40]. The bilayer structure is clearly seen, with the bcc structure at the
bottom and the hcp structure at the top. The diffraction pattern corresponding to this
HRTEM picture does indicate that the top hcp Co film has a two-domain structure
with c axes perpendicular to each other [38]. The thickness of the Co bcc structure is
found to increase with increasing the growth temperature in the grown temperature
range studied (60–160◦ C) [41].
A 1.4-nm bcc Co/GaAs (001) film, capped with 3.0-nm Au layer, measured by
ex situ MOKE at room temperature, is shown in Fig. 13. The film exhibits strong
uniaxial anisotropy with the easy axis along [110] direction. In order to extract
the fourfold anisotropy field, rotating MOKE (ROTMOKE) measurements for the
same sample [42] are shown in Fig. 13b. The ROTMOKE curve is fitted to the
equation [43]:

E/V = −μ0 MS H cos (θ − φ) + Ku sin2 φ + K1 /4sin2 (2φ + π/2) , (1)

where E is the free energy, V is the volume of the film, Ms is the saturation
magnetization, H is the magnetic field, θ is the angle between H and Co [110]
direction, and φ is the angle between magnetization and Co [110] direction. The

Fig. 12 Cross-sectional
HRTEM picture of the 29 nm
Co/GaAs (001) film prepared

at 150 C [40]. LICENSE #:
4813040909381
16 Metallic Magnetic Thin Films 823

Fig. 13 (a) Longitudinal


MOKE loops and (b)
ROTMOKE data of 1.4-nm
Co deposited at 120 ◦ C on
GaAs (001). The solid line is
the fit to Eq. (1) [41].
LICENSE #:
RNP/20/APR/024983

best fit yields the fourfold anisotropy field μ0 HK1 = 2 K1 /MS = −15.2 mT and
the uniaxial anisotropy field μ0 HKu = 2Ku /MS = 22.0 mT. Since the magnetic
moment of bcc Co is 1.53 μB [33], we deduce K1 = μ0 HK1 MS /2 = −6.5 kJ/m3 .
The negative K1 means that the easy axes of the fourfold anisotropy are along Co
<110> directions.
Furthermore, the anisotropy as a function of the growth temperature and
thickness was systematically investigated by ROTMOKE [41]. Figure 14 shows HK1
fitted from the ROTMOKE curves as a function of the Co film thickness for different
growth temperatures, 60, 120, and 160◦ C, and the thickness of the phase transition.
HK1 for Co thinner than 1.2 nm deposited at 60 ◦ C is negative, and the absolute
values increase monotonically with thickness, suggesting the bulk-like origin of the
fourfold anisotropy. When the Co is thicker than 1.2 nm, |HK1 | first decreases to zero
at 1.8 nm and then increases with increasing thickness (Fig. 14a). Above 1.2 nm, hcp
Co, which contributes a positive K1 [36] and starts to develop [38, 44], leading to the
increases of HK1 . For Co film grown at 120◦ C, K1 is negative for the bcc phase and
starts to increase when the hcp phase starts to develop. The bcc structure Co films
grown at 60–140◦ C always exhibit negative HK1 , as shown in Fig. 14d. Although
the bcc structure film grown at 160◦ C is obtained below 2 nm, no Kerr signal is
detected. It might be due to the strong reaction between Co and GaAs to form the
nonmagnetic bcc CoGaAs compound. Above 2 nm, HK1 or K1 is positive for the
mixture phases of the bcc and hcp structures (Fig. 14c).
824 D. Wu and X.-F. Jin

Fig. 14 HK1 as a function of Co thickness measured on the Co wedge samples grown at (a) 60 ◦ C,
(b) 120 ◦ C, and (c) 160 ◦ C. Each HK1 value is fitted from ROTMOKE experiments. The dashed
line denotes the structure phase transition thickness. (d) HK1 versus growth temperature. HK1 is
measured at the transition Co thickness. For Co grown at 160 ◦ C, there is no transition point [41].
The triangle data point is obtained from Refs. [38, 41]. LICENSE #: RNP/20/APR/024983

Ni/GaAs (001)

Unlike Fe and Co which have several structural phases in bulk, nickel only has
one bulk structure—fcc. It was reported that a very thin bcc Ni phase less than 6
atomic layers (or 3 unit cells) could be epitaxially grown on a Fe(001) substrate
[45, 46]. Above 6 ML, the structure is complicated [45]. The bcc structure Ni was
measured to be FM with a magnetic moment of 0.55–0.80 μB /atom [47, 48] or
0.4 ± 0.45 μB /atom by different groups [49]. However, TC was found to lower than
77 K in Ref. [45] and higher than 300 K in Ref. [49]. In the Ni/Fe (001) system,
the magnetic properties of bcc Ni can be strongly altered by Fe, considering the
strong hybridization and magnetic coupling at the Ni/Fe interface [45, 48–51] For
example, the enhanced Tc and induced magnetic moment of Ni could occur by the
strong interface coupling [52, 53]. Therefore, it is desirable to obtain bcc Ni on a
nonmagnetic substrate to investigate its intrinsic properties.
Theoretically, it was predicted that bcc Ni with a lattice constant of 0.2773 nm
at equilibrium is paramagnetic and a transition to an FM state occurs upon lattice
expansion beyond 0.2815 nm [54]. However, Guo and Wang found that the magnetic
phase transition occurs at 0.2730 nm [55].
We found that bcc Ni can be epitaxially grown on GaAs (001) [56, 57]. The
growth temperature of 170 K is critical to achieve the thickest bcc Ni films of
3.5 nm. Figure 15a shows a cross-sectional HRTEM image of Ni (3.5 nm)/GaAs
(001) with the electron beam along GaAs [110] direction. Ni is unambiguously
identified to have a bcc structure [40] with a lattice constant of 0.28 nm. The
orientation of the bcc Ni/GaAs (001) film is (001)[100]Ni //(001)[100]GaAs and
(001)[110]Ni //(001)[110]GaAs . XRD experiments confirm the structure to be bcc,
as shown in Fig. 15b. In addition to the GaAs (400) diffraction peak, a broad peak
16 Metallic Magnetic Thin Films 825

Fig. 15 (a) Fourier-filtered


HRTEM image of 3.5 nm bcc
Ni/GaAs (001) with the
electron beam along the [110]
direction. (b) Grazing angle
XRD spectrum with an

incident angle at 0.2 [57].
LICENSE #:
RNP/20/APR/024984

superimposed on top of the narrow GaAs (400) peak is clearly observed (inset of
Fig. 15b). This broad peak is identified to be the bcc Ni (200) diffraction from which
an in-plane lattice constant of 0.282 nm is readily derived. The absence of the fcc Ni
◦ ◦
diffraction peaks at ∼25.9 for (200) and ∼ 38.1 proves that the bcc phase is pure.
Figure 16a shows the MOKE signal at remanence measured from a Ni wedge
sample at 120 K as a function of the Ni film thickness [57]. The extrapolated dashed
line in Fig. 16a passes through the origin, indicating that there is no interfacial
dead layer. The vanishing MOKE signal below ∼1.2 nm is attributed to the reduced
Tc due to the finite-size scaling effect of Tc in ultrathin film [58]. In contrast, the
disappearance of the magnetic signal for the fcc Ni/Cu (001) film is around 0.6 nm
at 120 K [59]. To obtain Tc of bulk bcc Ni, Tc is fitted as a function of the bcc Ni
thickness (Fig. 16b) by the finite-size scaling law [58]:

T c (∞) − T c(l)/T c (∞) = (c/ l)λ , (2)

where Tc(∞) is Tc in the bulk limit, l is the film thickness, λ is a critical


exponent, and c is a parameter. The fitting yields Tc(∞) = 456 K, which is much
lower than Tc of bulk fcc Ni (627 K). The magnetic moment is measured to be
0.52 ± 0.08 μB /atom at 5 K by superconducting quantum interference device
(SQUID) magnetometry, in good agreement with the theoretically calculated value
of 0.54 μB /atom at the lattice constant of 0.282 nm [55].
The magnetocrystalline anisotropy K1 was determined by the ROTMOKE
technique [42, 43]. Figure 16c presents the ROTMOKE data as a function of the
in-plane magnetic field (μ0 H = 500 mT) direction for the 3.5-nm bcc Ni/GaAs
(001) film. Similar to the procedure of determining the anisotropies of epitaxially
grown Fex Cu1−x and Co films on GaAs (001) [41, 43], we fitted the experimental
curve by considering Ku and K1 . The excellent fitting yields (μ0 HKu = 230 mT
and HK1 = 180 mT (Fig. 16c). Using the foregoing obtained magnetic moment
of bcc Ni, we obtained K1 = 40 kJm−3 , which is comparable with that of bcc Fe
(K1 = 47 kJm−3 ). The positive sign of K1 indicates that bcc Ni has the same cubic
magnetic easy axis as bcc Fe but opposite to that of fcc Ni (K1 = −57 kJm−3 )
[60]. We note that the presence of the uniaxial anisotropy in the thick bcc Ni films
826 D. Wu and X.-F. Jin

Fig. 16 (a) MOKE signal at remanence as a function of the bcc Ni thickness measured at 120 K.
The inset shows a representative hysteresis loop. (b) Tc as a function of film thickness. The solid
line is a theoretical fit to Eq. (2) with Tc(∞) = 456 K, λ = 2.23, and c = 1.17 nm. (c) ROTMOKE
data as a function of the magnetization directions. The lines are the theoretical fitting with different
parameters [57]. LICENSE #: RNP/20/APR/024984

could suggest that the bcc lattice has some degree of in-plane shear strain, similar
to that of the thick bcc Fe layers on GaAs (001) [61]. In addition to the contrast in
the magnetic anisotropy between fcc Ni and bcc Ni, their electronic band structures
also show significant difference [57].

Structure and Magnetism of Magnetic 3d Transition Metal Alloys

Py/GaAs (001)

Permalloy (Py) usually refers to Fex Ni1−x alloys with a stoichiometry of


0.18 ≤ x ≤ 0.25. Py has a vanishingly small magnetocrystalline anisotropy
and magnetostriction but extremely large magnetic permeability. These unique
properties make Py one of the most important soft magnets for a variety of
applications, such as the cores of motors and the free layers of spin-valve magnetic
reading heads. However, it is still unclear from the fundamental point of view why
the cubic magnetic anisotropy energy of Py is so small, while both Fe and Ni have
quite large magnetic anisotropy energy [62], albeit of opposite sign. In addition to
the fundamental importance, a better understanding of the vanishing magnetic
anisotropy in Py is crucial for the rational design of innovative intermetallic
materials for various applications.
It has long been believed that the vanishing of K1 in Py stems from the
cancellation between the positive K1 of Fe and the negative K1 of Ni [62–64]. This
cancellation picture works surprisingly well for the interpretation of the magnetic
anisotropy phase diagram of Fex Ni1−x alloys [65, 66] and was even used to support
findings of some first-principles calculations [67, 68]. However, this idea has not
been experimentally verified.
16 Metallic Magnetic Thin Films 827

Fig. 17 GIXRD spectrum


with an incident angle at 0.3◦
for 5 nm Py, fitted by three
identified peaks described in
the text. The absence of peaks
at 25.9◦ and 38.1◦ indicates
the absence of fcc Py [70].
LICENSE #:
RNP/20/APR/024985

The structure of bulk Py is fcc. Since both Fe and Ni grow epitaxially on GaAs
(001) in the bcc structure [57, 69], it is possible to grow the bcc structure Fex Ni1−x
alloys on GaAs (001). The Fex Ni1−x alloys were grown at 200 K by co-deposition
using MBE. In addition to the RHEED and LEED patterns suggesting the successful
growth of the bcc Py film, similar to the bcc Fe and Ni films [56], the grazing
incident XRD (GIXRD) proves the bcc structure of the Py film directly. Figure 17
presents a GIXRD spectrum of the Au (2 nm)/Fe0.25 Ni0.75 (5 nm)/GaAs (001)
sample with a fixed photon energy at 8.052 keV and an incident angle at 0.3◦ [70].
The absence of peaks at 25.9◦ and 38.1◦ for fcc Py (200) and Py (220), respectively,
excludes the existence of the fcc phase. The peak at 33.1◦ is composed of three parts:
(1) an extremely narrow peak at 33.155◦ from the GaAs (400) diffraction (blue line
in the inset); (2) a broad peak from the bcc Py (200) diffraction with an in-plane
lattice constant of 0.2825 nm (red line); (3) a very broad and shallow peak from
the fcc Au (220) diffraction (green line). We found similar results for samples with
all compositions, meaning that bcc structure of Fex Ni1−x alloys is realized over the
whole range of stoichiometry.
To compare the magnetic properties of fcc and bcc Py, samples with two
structures were prepared on one GaAs (001) substrate, as shown in the inset of
Fig. 18. Half of the substrate was first covered by a 3-nm fcc Au seed layer by
a shadow mask [71]. After removing the shadow mask, a thickness wedge of Py
was then deposited on the whole sample. Therefore, we had fcc Py/Au/GaAs (001)
[71] and bcc Py/GaAs (001) films with exactly the same stoichiometry on the same
sample.
Figure 18 shows the Kerr rotations at saturation for two structural Py as a function
of the film thickness measured at 110 K. The different slope of Kerr rotations are
attributed to the optical effect because the magnetic moment per atom is almost
identical for bcc and fcc Py: mbcc = 1.03 ± 0.09μB and mfcc = 1.07 ± 0.09μB
measured on the corresponding flat film by vibrating sample magnetometry (VSM).
828 D. Wu and X.-F. Jin

Fig. 18 Kerr rotations of bcc


and fcc Fe0.25 Ni0.75 as a
function of the film thickness.
The right inset shows the
special sample prepared for
this measurement. The left
inset gives the finite-size
scaling of Tc of the bcc Py
film; dots for experimental
data and line for the fitting
[70]. LICENSE #:
RNP/20/APR/024985

Furthermore, Tc as a function of the film thickness in a bcc Py wedge was measured


and shown in the inset of Fig. 18. Tc of the bulk bcc Py is obtained to be 553 K by
fitting the curve to Eq. (2). It is significantly lower than Tc = 871 K for the bulk fcc
Py [72].
The sample was covered by 2-nm-thick Au to perform ex situ longitudinal
MOKE measurements. The top left inset of Fig. 19 shows the typical hysteresis
loops of the 3-nm-thick bcc Py. Except a small uniaxial magnetic anisotropy caused
by the substrate as generally found for 3d magnetic metals on III–V semiconductors
[73], the in-plane cubic magnetic anisotropy is extremely small. To understand
the anisotropy of Py, four compositional wedges of the bcc Fex Ni1−x /GaAs (001)
samples were fabricated to cover the entire stoichiometry range [74]. K1 was
obtained by the ROTMOKE measurements at room temperature for the whole range
of compositions, as shown in Fig. 19 [42, 43]. K1 (x) curve shows a tubelike shape.
In contrast, K1 (x) curve of fcc Fex Ni1−x is positive at Fe-rich side, passes through
zero at x = 25%, and becomes negative at the Ni-rich side.
The magnetization versus concentration of bcc Fex Ni1−x measured by VSM on
the flat sample obeys very well the Slater-Pauling behavior, as shown in the top
right inset of Fig. 19, in comparison to the high-spin to low-spin transition in the fcc
Fex Ni1−x [75]. Moreover, it is observed in the same inset that the magnetization
of bcc Fex Ni1−x measured by MOKE has nothing peculiar nearby the eminent
“Invar point” (65% Fe), where the fcc-bcc structural phase transition happens in
the conventional bulk Fex Ni1−x .
First-principles calculations using the full potential linearized augmented plane
wave method in conjunction with the generalized gradient approximation (GGA)
clarify the experimental results [76, 77]. When x = 0.25, with the (2 × 2 × 2)
supercells shown in the inset of Fig. 18b, the calculated spin magnetic moments
for bcc and fcc Py are mbcc = 1.11μB /atom and mfcc = 1.14μB /atom, respectively.
These results are in good accordance with the VSM data. Since the calculations
were done in a chemically ordered supercell (e.g., the DO3 structure for the bcc
lattice), the chemical disordering in experimental samples appears to be unimportant
16 Metallic Magnetic Thin Films 829

Fig. 19 Cubic magnetic anisotropy constant (K1 ) versus iron concentration measured from the
3-nm-thick bcc Fex Ni1−x composition wedge (see the carton) by ROTMOKE at 300 K, together
with the calculated K1 values are marked as red solid cycles. A set of representative hysteresis loops
for bcc Fe0.25 Ni0.75 is shown in the inset. Averaged magnetic moment per atom in bcc Fex Ni1−x is
also measured by VSM as a function of iron concentration together with a zoom around the “Invar
point” obtained by MOKE [70]. LICENSE #: RNP/20/APR/024985

for the magnetization [78]. The cubic magnetocrystalline anisotropy energies were
calculated through the torque method for bcc Fex Ni1−x in several (2 × 2 × 2)
supercells that preserve the cubic symmetry [79]. The results are presented as
dots in Fig. 19 for direct comparison with the experiment. It is obvious that the
theory reproduces well the general trend of the experimental data. The band-filling
dependence of K1 obtained from a rigid band scheme is presented in Fig. 20 for
several systems. It is clear to see that adding either a small amount of Fe to Ni
or vice versa causes a reduction in K1 as guided by the arrows. To elucidate the
element-specific contributions from Fe and Ni, we selectively switch off spin-orbit
coupling (SOC) from different elements in the bcc Py. Intriguingly, the SOC of
Fe only (denoted as Fe-SOC) gives a small negative contribution to K1 whereas
Ni plays the dominant role. Both Ni and Fe behave very differently from their bulk
forms. It seems that only the local Fe-Ni bonding is the most important factor for the
quenching of the magnitude of K1 , whereas signs of K1 of individual constituents
are not relevant.

Cox Mn1−x /GaAs (001)

Among the 3d transition metal alloys , the Mn-based alloys have attracted special
attention because of their potential applications in the magnetic recording industry
[80–82] The Cox Mn1−x alloy has shown interesting magnetic properties [83–85]
Bulk Cox Mn1−x is FM with an hcp structure for 0 < x ≤ 0.32 but is AFM with an
fcc structure for 0.32 < x ≤ 0.52. For x > 0.52; it is no longer possible to stabilize
the fcc state [86]. The metastable bcc Co and fcc Mn can be epitaxially grown on
GaAs (001) surfaces at room temperature [13, 38]. It provides us an ideal platform to
830 D. Wu and X.-F. Jin

Fig. 20 Calculated K1 of Fe,


Ni, and Fe3 Ni13 as a function
of band filling. The vertical
line at zero corresponds to the
Fermi energy. “Fe-SOC”
means only spin-orbit
coupling of Fe is included and
vice versa [70]. LICENSE #:
RNP/20/APR/024985

epitaxially grow the Cox Mn1−x alloys on GaAs (001) to investigate the correlation
between their structure and magnetism.
Figure 21 presents the typical RHEED patterns of the Cox Mn1−x /GaAs (001)
films with different compositions with the incident electron beam along the [110]
direction [87]. The spot-like patterns indicate island growth mode and allow us
to determine the structure [14]. Figure 21a shows the RHEED pattern for a clean
GaAs (001) surface, which serves as an internal scale to measure the film lattice
constants. Figure 21b shows a typical RHEED pattern for the Cox Mn1−x films with
composition 0 < x ≤ 0.44. The rectangular-shaped pattern corresponds to a bcc
structure, which does not exist in the bulk and is similar to that of the Co/GaAs (001)
film. The in-plane lattice constants of Co60 Mn40 are estimated to be a = b = 0.29 nm
from the RHEED pattern (Fig. 21b). The out-of-plane lattice constant is determined
to be 0.289 nm by XRD.
For 0.78 ≤ x < 1, the RHEED pattern is square, as shown in Fig. 21c, suggesting
an fcc structure, similar to that of the Mn/GaAs (001) film. In comparison, no stable
bulk phases of Cox Mn1−x alloy can be obtained for x > 0.52. Similar to the above
analyses, the lattice constants are estimated to be a = b = 0.36 nm and c = 0.360 nm
by the RHEED pattern and XRD.
The structure is more complicated and shows thickness dependence for
0.44 ≤ x ≤ 0.78. A typical case is given in Fig. 21d–f, in which the RHEED
patterns of Co30 Mn70 thin films were recorded at different thicknesses. In the initial
stage at film thickness thinner than 5 nm, the film is bcc (Fig. 21d). As the film gets
thicker, a mixture of rectangular- and square-like diffraction patterns is observed
(Fig. 21e), indicating the coexistence of the fcc and bcc phases on top of the initial
bcc layer. When the film is thicker than ∼27 nm, the film becomes fcc (Fig. 21f).
The coexistence of the fcc and bcc phases is further confirmed by XRD [87]. The
thickness-dependent behavior in the intermediate range of composition is due to the
competition between the fcc structure as Co/GaAs (001) film and bcc structure as
Mn/GaAs (001) film. In the intermediate composition range, the bcc phase of the
Cox Mn1−x alloys has a much smaller lattice misfit with the substrate (∼ 1%) than
16 Metallic Magnetic Thin Films 831

Fig. 21 Representative RHEED patterns of Cox Mn1−x /GaAs (001) with different Mn composi-
tion. (a) Clean GaAs (001) surface; (b) bcc phase of Co60 Mn40 films; (c) fcc phase of Co10 Mn90
films; (d–f) thickness-dependent transition from bcc to fcc phase for the Co30 Mn70 film [87].
LICENSE #: RNP/20/APR/024986

the fcc phase (∼ 9%), resulting in the initial growth of the bcc phase. However, the
bcc phase does not exist in the bulk, indicating a relatively high-energy metastable
state. When the film gets thicker, it tends to form the fcc phase. It is very unlikely
that the bcc and fcc coexistence region corresponds to a phase separation of Mn-rich
fcc and Co-rich bcc phases; otherwise the film structure in the final stage of growth
would become much more complicated than the one we observed (Fig. 21f).
Based on the foregoing results, a structural phase diagram as a function of
composition is plotted in Fig. 22. Three regions are clearly seen, i.e., the Co-rich
bcc phase region marked I, the Mn-rich fcc phase region marked III, and the
intermediate region with the mixed bcc and fcc structures marked II.
832 D. Wu and X.-F. Jin

Fig. 22 Structural phase diagram for Cox Mn1−x /GaAs (001) films [87]. LICENSE #:
RNP/20/APR/024986

After the structure characterization, the Cox Mn1−x /GaAs (001) films were
carried out with the MOKE measurements. The films are in-plane anisotropy
if they are FM. Figure 23 shows the typical longitudinal MOKE loops for the
Cox Mn1−x /GaAs (001) films with different compositions after being normalized
for the film thickness. The film shows strong ferromagnetism for x < 0.44. No
hysteresis loop is observed for x > 0.78. The ferromagnetism is relatively weak
for 0.44 ≤ x ≤ 0.78.
Since the structure of the Cox Mn1−x /GaAs (001) films is thickness-dependent in
the intermediate composition, a wedge-shaped sample was fabricated and measured
by MOKE at room temperature to understand the correlation between the structure
and magnetism. The results of the Kerr intensity versus thickness for a Co52 Mn48
film are shown in Fig. 24. The data can be divided into three regions: i) a magnetic
dead layer (∼ 2 nm) is near the interface; ii) the linearly increasing Kerr intensity
(∼ 2–6 nm) region means an FM layer; iii) the saturated Kerr intensity region
(> 6 nm) means a non-FM layer on the top. The magnetic dead layer is caused by the
interface reaction between film and substrate, as is quite common in such systems.
The FM layer comes from the bcc-dominated Co52 Mn48 film. The non-FM layer
has an fcc-dominated structure. Furthermore, the fcc-Cox Mn1−x /Fe bilayer exhibits
an exchange bias effect, indicating the fcc structure Cox Mn1−x is AFM. Therefore,
we conclude that Cox Mn1−x film is FM whenever its structure is bcc, and it is AFM
whenever its structure is fcc. The magnetism is intimately correlated with a structure
16 Metallic Magnetic Thin Films 833

Fig. 23 Longitudinal MOKE


hysteresis curves for the
Cox Mn1−x /GaAs (001) films
at different Mn compositions
[87]. LICENSE #:
RNP/20/APR/024986

for the Cox Mn1−x alloys. This strong correlation realized in experiments between
the structure and magnetism of Cox Mn1−x alloys was further studied and confirmed
by a first-principles linearized augmented plane-wave (LAPW) calculation with the
local-spin-density approximation (LSDA) [87]. This correlation between structure
and magnetism is also found in Fex Mn1−x /GaAs (001) films [88].

Fex Cu1−x /GaAs (001)

Since bulk Fe has the bcc structure, the magnetic properties of fcc Fe attract
tremendous interest. Fcc Fe can only be prepared via precipitation in a Cu matrix
or epitaxially grown on a Cu substrate [31]. Because of the extremely low bulk
miscibility of either Cu in Fe or Fe in Cu, only ∼4% Fe dissolves into Cu (with
fcc structure) and ∼ 10% Cu into Fe (with bcc structure) near their respective
liquid points [89], and a long-standing challenge has been the goal of producing
metastable Fex Cu1−x alloys with higher concentrations by different approaches,
such as rapid quenching, vapor deposition, ion beam mixing, or mechanical alloying
[90–94]. However, the samples prepared by these methods are polycrystalline. A
834 D. Wu and X.-F. Jin

Fig. 24 Kerr intensity for a wedge-shaped Co52 Mn48 film at different thicknesses [87]. LICENSE
#: RNP/20/APR/024986

monocrystalline sample was achieved in a monolayer stacked Fe/Cu superlattice


grown on the Cu (001) substrate by pulsed laser deposition [95]. This superlattice
corresponds to only one composition of x = 0.5, Fe0.5 Cu0.5 .
The magnetic properties of the Fex Cu1−x alloys with increasing x is contro-
versial. It was reported that the Fe moment remains ∼2.2 μB /atom over a wide
composition range at x > 0.50 and then quickly falls to zero as x decreases
experimentally [96]. However, another group found that there were no noticeable
changes in the magnetic moment over the entire composition range [91]. First-
principles calculations showed that the magnetic moment per Fe atom was almost
independent of Fe-content over a wide range (x > 0.5) in both the bcc and fcc phases
and then decreased and reached zero for x ≤ 0.25 [97]. However, it was found that
the magnetic moment per Fe atom in the bcc phase increased from 2.2 μB to 2.62
μB as x decreased from 1 to 0.5 after taking into account the magneto-volume effect
in the calculation [98].
Fe can be epitaxially grown on GaAs (001) with the bcc structure [39]. The
RHEED pattern of bcc Fe/GaAs (001) film is shown in Fig. 25a [43]. The aspect
ratio of a*/b*, shown in Fig. 25a, is close to 1.4. The RHEED patterns of the
Cu/GaAs (001) films transform from rectangular to square shape as increasing the
film thickness, as shown in Fig. 25b–d, meaning that the film is almost bcc-like
for the first several layers, then body-centered tetragonal (bct), and finally in its
thermodynamically stable fcc structure. The evolution of the corresponding aspect
ratio of a*/b* extracted from the RHEED patterns as a function of the film thickness
is plotted in Fig. 26.
16 Metallic Magnetic Thin Films 835

Fig. 25 RHEED patterns for


(a) Fe film, (b–d) Cu film,
and (e–h) Fe0.25 Cu0.75 film
on GaAs (001) with the
electron beam along the [110]
direction [43]. LICENSE #:
RNP/20/APR/024987

Fig. 26 The a*/b* aspect


ratio extracted from RHEED
pattern as a function of the
film thickness for Fex Cu1−x
at different x [43]. LICENSE
#: RNP/20/APR/024987

Similar to Cu/GaAs (001) film, the structure varies with the film thickness for
the Fex Cu1−x films with any particular x value. Figure 25e–h shows a representative
example of Fe0.25 Cu0.75 . The structure evolves from bcc to bct (closer to bcc), then
to bct but closer to fcc, and finally to fcc. The structural evolution for different
Fex Cu1−x films is summarized by the aspect ratio of a*/b* in Fig. 26. At the
same thickness of 6 nm, a*/b* changes from 1 (fcc) for Cu to 1.4 (bcc) for Fe
as increasing x.
The Fex Cu1−x /GaAs (001) films were covered by 2 nm Au to conduct ex situ
MOKE measurements. It is found that the ferromagnetic to non-ferromagnetic
836 D. Wu and X.-F. Jin

Fig. 27 (a) The Kerr


intensity at saturation is
plotted as a function of the
film thickness for Fe0.5 Cu0.5 .
The inset shows a typical
MOKE hysteresis loop, from
which the Kerr intensity at
saturation is obtained; (b) the
effective magnetic moment
per Fe concentration [43].
LICENSE #:
RNP/20/APR/024987

transition occurs at x ≈ 0.33. This result is quite different from the polycrystalline
samples, where the transition is at x ≈ 0.5 [97].
Considering that the structure of the Fex Cu1−x films is dependent on the film
thickness, the magnetization for different thickness was measured in the wedge
samples at room temperature. The result of Fe0.5 Cu0.5 is shown in Fig. 27a. The
straight line above 2 nm across origin indicates that there is no noticeable change of
effective magnetic moment per Fe atom. The deviation from the straight line below
2 nm is due to the finite size effect of Tc [58]. The SQUID measurements were
carried out at 5 K on the 6-nm Fex Cu1−x films to quantify the effective magnetic
moment per Fe atom, assuming that Cu does not contribute to the magnetization.
The results are shown in Fig. 27b. The effective magnetic moment per Fe atom
decreases with increasing x, consistent with the previous theoretical calculation
which is also plotted in the figure [98]. For the 6-nm-thick Fe0.33 Cu0.67 , the
magnetization is too small to be detected.
It is known that Fe/GaAs (001) has a fourfold anisotropy with the easy axis along
the <100> directions and a uniaxial anisotropy with the easy axis along the [110]
direction [69]. The total energy density of the system is

E/V = −μ0 MS H cos (α − θ ) + Ku sin2 θ


 ◦
  ◦
 (3)
+ K1 sin2 θ + 45 + K1 cos2 θ + 45 .

Here, the first term represents the Zeeman energy, α is the angle between the
easiest axis (ea) and H (see the inset of Fig. 28b), and θ is the angle between the
easiest axis and the magnetization; the second term is the uniaxial anisotropy energy,
and the third term is the fourfold anisotropy energy. In order to extract the anisotropy
field, we carried out the ROTMOKE measurements. The results are shown in
Fig. 28. The ROTMOKE curve (Fig. 28a) is fitted to Eq. (3) (solid line in Fig. 28a) to
yield the anisotropy field μ0 HKu = Ku /MS = 6.2 mT and μ0 HK1 = K1 /MS = 60.4
mT, respectively, for the Fe/GaAs (001) film. For the Fe0.50 Cu0.50 /GaAs (001) film,
16 Metallic Magnetic Thin Films 837

Fig. 28 The torque


momentl(θ) as a function of
θ, (a) μ0 H = 80 mT, 4 nm
Fe; (b) μ0 H = 20 mT, 8 nm
Fe0.50 Cu0.50 . A schematic
picture of the applied field H
and the magnetization M is
shown in the inset [43].
LICENSE #:
RNP/20/APR/024987

it is obvious that the ROTMOKE curve (Fig. 28b) cannot be fitted by considering
only the uniaxial anisotropy (dashed line in Fig. 28b). The curve is well fitted with
both uniaxial and fourfold anisotropies for μ0 HKu = 6.9 mT and μ0 HK1 = 4.7
mT as shown by the solid line in Fig. 28b. Although the fourfold anisotropy of
Fe0.50 Cu0.50 /GaAs (001) is much weaker than that of pure Fe, it still retains.

Fex Pd1−x /Cu (100)

It is well-known that the magnetic long-range order does not exist in an isotropic
two-dimensional (2D) Heisenberg system at any finite temperature [99]. However, it
is hard to physically realize a 2D FM film without any anisotropy. Experimentally, it
might be realized in the vicinity of the spin reorientation transition (SRT) in ultrathin
magnetic films, because of the magnetocrystalline anisotropy, surface anisotropy,
and dipolar anisotropy could compensate at the SRT point in some cases. Therefore,
it is very interesting to study the correlation between SRT and the long-range
magnetic order. Indeed, a suppression of magnetization or a reduction of Tc within
the SRT region was reported in Fe/Cu (001), Fe/Ag (001), and Fe/Ni/Cu (001) [100–
102]. However, if it is a continuous SRT, the anisotropy never disappears within
the SRT region, where the magnetization only rotates between in-plane and out-of-
plane as Ni/Cu (001) [103, 104]. Interestingly, it was found that the SRT occurs at
87% Ni concentration in 9-ML Fex Pd1−x /Cu (100) and Tc is dramatically reduced
[105]. Therefore, the Fex Pd1−x alloy provides us a good system to investigate the
correlation between SRT and Tc [106].
A Fex Pd1−x composition wedge sample was epitaxially grown on Cu (001)
to explore the composition-dependent SRT [74]. The magnetic properties of the
sample were investigated through both MOKE and ac susceptibility measurements
after capping 3 ML Cu on the sample [52, 107–109]. Two representative hysteresis
loops of 9-ML sample measured by MOKE at room temperature were shown in
838 D. Wu and X.-F. Jin

Fig. 29 (a) The longitudinal


MOKE hysteresis loop of
Ni0.57 Pd0.43 and polar MOKE
hysteresis loop of
Ni0.91 Pd0.09 measured on a
9-ML composition wedge
sample. (b) The imaginary
part of the ac susceptibility
Im(χ) and remanent
magnetization determined
from polar MOKE of 9-ML
composition wedge sample as
a function of Ni composition
[106]. LICENSE #:
RNP/20/APR/024988

Fig. 29a, one for Ni0.57 Pd0.43 with easy axis in-plane and the other for Ni0.91 Pd0.19
with a perpendicular easy axis. The remanent magnetization measured by the polar
MOKE as a function of the Ni composition is plotted in Fig. 29b. The remanent
magnetization increases gradually with increasing Ni concentration, suggesting that
the magnetization rotates continuously from in-plane to out-of-plane. The imaginary
part of the ac susceptibility Im(χ ) as a function of the Ni composition was also
shown in Fig. 29b. Im(χ ) shows a pronounced peak at the SRT position, which
is at 86% Ni concentration for 9-ML Fex Pd1−x /Cu (100), consistent with the
previous report [105]. This SRT can be understood by the competition between the
shape anisotropy and the magnetocrystalline anisotropy including magnetoelastic
anisotropy originating from the lattice distortion and the surface anisotropy, similar
to the Ni/Cu (001) system [104, 110, 111].
Tc of the 9-ML compositional Fex Pd1−x /Cu (100) film is determined by the
ac susceptibility measurement, as shown in an example in the inset of Fig. 30a.
The result of Tc and the susceptibility as a function of Ni composition is shown
in Fig. 30a. The reduction of Tc is not found at the SRT position where the
susceptibility exhibits a peak. Tc increases very slowly as a function of the Ni
composition before SRT, while it increases more sharply after SRT. This behavior of
the turning point of Tc at SRT indicates that the variation in anisotropy influences
the increasing behavior of Tc. To confirm the correlation between SRT and Tc, a
7-ML compositional Fex Pd1−x /Cu (100) film was fabricated and measured using
16 Metallic Magnetic Thin Films 839

Fig. 30 (a) Im(χ) and Curie


temperatures versus Ni
composition of 9-ML
Nix Pd1−x composition wedge
sample. An example of Curie
temperature measurement by
ac susceptibility is given in
the inset, where the remanent
magnetization determined by
MOKE as a function of
temperature is also shown.
(b) Im(χ) and Curie
temperatures versus Ni
composition for 7 ML
Nix Pd1−x composition wedge
sample [106]. LICENSE #:
RNP/20/APR/024988

Fig. 31 Curie temperatures


versus increasing
perpendicular anisotropy and
fourfold in-plane anisotropy,
in which K is the anisotropy
parameter and J is the FM
exchange constant [106].
LICENSE #:
RNP/20/APR/024988

the same approach. The results are shown in Fig. 30b. Again, the turning point of
Tc is at the SRT point, which is at a Ni concentration of 96% and no reduction in
Tc is observed in the region of SRT, unlike another report (Ref. [105]).
No reduction in Tc is observed around the SRT region in our experiments. One
reason could be that its anisotropy never really disappears, but the easy direction
of the magnetization rotates continuously within the transition [112]. The other
possibility could be that at the SRT point there still exists some residual perpendic-
ular anisotropy which is caused by variations in the perpendicular anisotropy across
the film surface [113]. Another interesting phenomenon in our experiments is that
increasing the perpendicular anisotropy may enhance Tc much more strongly than
the fourfold in-plane anisotropy. A simple Monte Carlo simulation was performed
840 D. Wu and X.-F. Jin

to calculate Tc of a 2D Heisenberg model with perpendicular anisotropy and


fourfold in-plane anisotropy, respectively. The calculation shows that Tc increases
much faster by increasing the perpendicular anisotropy than the fourfold anisotropy,
shown in Fig. 31. This is because the fourfold in-plane anisotropy is a higher-
order term compared to the perpendicular uniaxial anisotropy. The perpendicular
anisotropy breaks the symmetry of the system more seriously than the fourfold one.
As a consequence, uniaxial anisotropy should be much more effective at triggering
ferromagnetism than fourfold anisotropy in a low-dimensional system.

Acknowledgments One of the authors, XFJ, acknowledges the invaluable contribution of the past
co-workers of Prof. G. S. Dong and graduated students.

References
1. Baibich, M.N., Broto, J.M., Fert, A., Nguyen Van Dau, F., Petroff, F., Eitenne, P., Creuzet,
G., Friederich, A., Chazelas, J.: Giant magnetoresistance of (001) Fe/(001) Cr magnetic
superlattices. Phys. Rev. Lett. 61, 2472 (1988)
2. Binash, G., Grünberg, G.P., Saurenbach, F., Zinn, W.: Enhanced magnetoresistance in layered
magnetic structures with antiferromagnetic interlayer exchange. Phys. Rev. B. 39, 4828
(1989)
3. Pauling, L.: The nature of the interatomic forces in metals. Phys. Rev. 54, 899–904 (1938)
4. Slater, J.C.: Electronic structure of alloys. J. Appl. Phys. 8, 385–390 (1937)
5. Miron, I.M., Garello, K., Gaudin, G., Zermatten, P.-J., Costache, M.V., Auffret, S., Bandiera,
S., Rodmacq, B., Schuhl, A., Gambardella, P.: Perpendicular switching of a single ferromag-
netic layer induced by in-plane current injection. Nature. 2011(476), 189 (2011)
6. Jin, X.F.: Interfaces between magnetic thin films and GaAs substrate. J. Electr. Spectr. Relat.
Phen. 114–116, 771 (2001)
7. Qian, D., Liu, G.L., Loison, D., Dong, G.S., Jin, X.F.: Growth and structure of Cr thin films
on GaAs (0 0 1). J. Cryst. Grow. 218, 197 (2000)
8. Wolf, J.A., Krebs, J.J., Prinz, G.A.: Growth and magnetic characterization of face centered
cubic Co on (001) diamond. Appl. Phys. Lett. 65, 1057 (1994)
9. Weiss, R.J., Tauerr, K.J.: Thermodynamics and magnetic structures of the allotropic modifi-
cations of manganese. J. Phys. Chem. Solids. 4, 135 (1958)
10. Tian, D., Li, H., Wu, S.C., Jona, F., Marcus, P.M.: Epitaxy of Mn on Pd (001). Solid State
Commun. 70, 199 (1989)
11. Jonker, B.T., Krebs, J.J., Prinze, G.A.: Growth and magnetic characterization of Mn films and
superlattices on Ag (001). Phys. Rev. B. 39, 1399 (1989)
12. Idzerda, Y.U., Jonker, B.T., Elam, W.T., Prinze, G.A.: Identification of the Mn structure in
Mn/Ag (001) superlattices. J. Appl. Phys. 67, 5385 (1990)
13. Jin, X., Zhang, M., Dong, G.S., Xu, M., Chen, Y., Xun Wang, X.G.Z., Shen, X.L.:
Stabilization of face-centered-cubic Mn films via epitaxial growth on GaAs (001). Appl. Phys.
Lett. 65, 3078 (1994)
14. Jin, X., Chen, Y., Lin, X.W., Dong, D.S., Chen, Y., Xu, M., Zhu, W.R., Wang, X., Shen, X.L.,
Li, L.: Interface structure of fcc Mn on GaAs (001). Appl. Phys. Lett. 70, 2455 (1997)
15. Dong, G.S., Xu, M., Chen, Y., Jin, X., Xun Wang: XPS study of Mn thin films grown on GaAs
(001) surfaces. Surf. Interface Anal. 24, 653 (1996)
16. Wang, C.S., Klein, B.M., Krakauer, H.: Theory of magnetic and structural ordering in iron.
Phys. Rev. Lett. 54, 1852 (1985)
17. Pinski, F.J., Staunton, J., Gyorffy, B.L., Johnson, D.D., Stocks, G.M.: Ferromagnetism versus
antiferromagnetism in face-centered-cubic iron. Phys. Rev. Lett. 56, 2096 (1986)
16 Metallic Magnetic Thin Films 841

18. Moruzzi, V.L., Marcus, P.M., Schwarz, K., Mohn, P.: Ferromagnetic phases of bcc and fcc
Fe, Co, and Ni. Phys. Rev. B. 34, 1784 (1986)
19. Myrasov, O.N., Gubanov, V.A., Liechtenstein, A.I.: Spiral-spin-density-wave states in
fcc iron: Linear-muffin-tin-orbitals band-structure approach. Phys. Rev. B. 45, 12330
(1992)
20. Uhl, M., Sandratskii, L.M., Kubler, J.: Spin fluctuations in γ-Fe and in Fe3 Pt Invar from
local-density-functional calculations. Phys. Rev. B. 50, 291 (1994)
21. Körling, M., Ergon, J.: Gradient-corrected ab initio calculations of spin-spiral states in fcc-Fe
and the effects of the atomic-spheres approximation. Phys. Rev. B. 54, R8293 (1996)
22. Zhou, Y.M., Wang, D.S., Kawazoe, Y.: Effective ab initio exchange integrals and magnetic
phase transition in fcc Fe and Mn antiferromagnets. Phys. Rev. B. 59, 8387 (1999)
23. Tsunoda, Y.: Spin-density wave in cubic γ-Fe and γ-Fe100-x Cox precipitates in Cu. J. Physica.
C. 1(10), 427 (1989)
24. Tsunoda, Y., Nishioka, Y., Nicklow, R.M.: Spin fluctuations in small γ-Fe precipitates. J.
Magn. Magn. Mater. 128, 133 (1993)
25. Kalki, K., Chambliss, D.D., Johnson, K.E., Wilson, R.J., Chiang, S.: Evidence for martensitic
fcc-bcc transition of thin Fe films on Cu (100). Phys. Rev. B. 48, 18344 (1993)
26. Wuttig, M., Feldmann, B., Thomassen, J., May, F., Zillgen, H., Brodde, A., Hannemann, H.,
Neddermeyer, H.: Structural transformations of fcc iron films on Cu (100). Surf. Sci. 291, 14
(1993)
27. Zillgen, H., Feldmann, B., Wuttig, M.: Structural and magnetic properties of ultrathin Fe films
deposited at low temperature on Cu (100). Surf. Sci. 321, 32 (1994)
28. Thomassen, J., Feldmann, B., Wuttig, M., Ibach, H.: Magnetic live surface layers in Fe/Cu
(100). Phys. Rev. Lett. 69, 3831 (1992)
29. Keune, W., Ezawa, T., Macedo, W.A.A., Glos, U., Schletz, K.P., Kirschbaum, U.: Magneto-
volume effects in γ-Fe ultrathin films and small particles. Physica (Amsterdam). 161B, 269
(1989)
30. Li, D., Freitag, M., Pearson, J., Qiu, Z.Q., Bader, S.D.: Magnetic and structural instabilities
of ferromagnetic and antiferromagnetic Fe/Cu (100). J. Appl. Phys. 76, 6425 (1994).; Phys.
Rev. Lett. 72, 3112 (1994)
31. Qian, D., Jin, X.F., Barthel, J., Klaua, M., Kirschner, J.: Spin-density wave in ultrathin Fe
films on Cu (100). Phys. Rev. Lett. 87, 227204 (2001)
32. Shen, J., Schmidthals, C., Woltersdorf, J., Kirschner, J.: Structural phase transformation under
reversed strain: a comparative study of iron ultrathin film growth on nickel and copper (100).
Surf. Sci. 407, 90 (1998)
33. Prinz, G.A.: Stabilization of bcc Co via epitaxial growth on GaAs. Phys. Rev. Lett. 54, 1051
(1985)
34. Liu, A.Y., Singh, D.J.: Elastic instability of bcc cobalt. Phys. Rev. B. 47, 8515 (1993)
35. Blundell, S.J., Gester, M., Bland, J.A.C., Daboo, C., Gu, E., Baird, M.J., Ives, A.J.R.:
Structure induced magnetic anisotropy behavior in Co/GaAs (001) films. J. Appl. Phys. 73,
5948 (1993)
36. Gu, E., Gester, M., Hicken, R.J., Daboo, C., Tselepi, M., Gray, S.J., Bland, J.A.C., Brown,
L.M.: Fourfold anisotropy and structural behavior of epitaxial hcp Co/GaAs (001) thin films.
Phys. Rev. B. 52, 14704 (1995)
37. Klautau, A.B., Eriksson, O.: Ab initio calculation of the magnetocrystalline anisotropy and
spin and orbital moments of a bcc Co (001) surface. Phys. Rev. B. 72, 014459 (2005)
38. Wu, Y.Z., Ding, H.F., Jing, C., Wu, D., Liu, G.L., Gordon, V., Dong, G.S., Jin, X.F., Zhu, S.,
Sun, K.: In-plane magnetic anisotropy of bcc Co on GaAs (001). Phys. Rev. B. 57, 11935
(1998)
39. Waldrop, J.R., Grant, R.W.: Interface chemistry of metal-GaAs Schottky-barrier contacts.
Appl. Phys. Lett. 34, 630 (1979).; Prinz, G.A., Krebs, J.J.: Molecular beam epitaxial growth
of single-crystal Fe films on GaAs. ibid. 39, 397 (1981); Dong, G.S., Chen, Y., Zhang, M., Jin,
X.: Simple design of electron beam evaporators for 3d transition metals. J. Vac. Sci. Technol.
A 13, 159 (1995)
842 D. Wu and X.-F. Jin

40. Wu, Y.Z., Ding, H.F., Jing, C., Wu, D., Dong, G.S., Jin, X.F., Sun, K., Zhu, S.: Epitaxy and
magnetism of Co on GaAs (001). J. Mag. Mag. Mat. 198-199, 297 (1999)
41. Xu, X.Y., Yin, L.F., Wei, D.H., Tian, C.S., Dong, G.S., Jin, X.F., Jia, Q.J.: Measurement of
the thickness-dependent magnetic anisotropy of Co/GaAs (001). Phys. Rev. B. 77, 052403
(2008)
42. Mattheis, R., Quednau, G.: Determination of the anisotropy field strength in ultra-thin
magnetic films using longitudinal MOKE and a rotating field: the ROTMOKE method. J.
Magn. Magn. Mater. 205, 143 (1999)
43. Tian, Z., Tian, C., Yin, L., Wu, D., Dong, G., Jin, X., Qiu, Z.: Magnetic ordering and
anisotropy of epitaxially grown Fex Cu1-x alloy on GaAs (001). Phys. Rev. B. 70, 012301
(2004)
44. Mangan, M.A., Spanos, G., Ambrose, T., Prinz, G.A.: Transmission electron microscopy
investigation of Co thin films on GaAs (001). Appl. Phys. Lett. 75, 346 (1999)
45. Heinrich, B., Purcell, S.T., Dutcher, J.R., Urquhart, K.B., Cochran, J.F., Arrott, A.S.:
Structural and magnetic properties of ultrathin Ni/Fe bilayers grown epitaxially on Ag (001).
Phys. Rev. B: Cond. Matter. 38, 12879 (1988)
46. Wang, Z.Q., Li, Y.S., Jona, F., Marcus, P.M.: Epitaxial growth of body-centered-cubic nickel
on iron. Solid State Commun. 61, 623 (1987)
47. Bland, J.A.C., Bateson, R.D., Johnson, A.D., Heinrich, B., Celinski, Z., Lauter, H.J.:
Magnetic properties of ultrathin bcc Fe (001) films grown epitaxially on Ag (001) substrates.
J. Magn. Magn. Mater. 93, 331 (1991)
48. Celinski, Z., Urquhart, K.B., Heinrich, B.: Using ferromagnetic resonance to measure the
magnetic moments of ultrathin films. J. Magn. Magn. Mater. 166, 6 (1997)
49. Brookes, N.B., Clarke, A., Johnson, P.D.: Electronic and magnetic structure of bcc nickel.
Phys. Rev. B. 46, 237 (1992)
50. Lee, J.I., Hong, S.C., Freeman, A.J., Fu, C.L.: Enhanced surface and interface magnetism of
bcc Ni overlayers on Fe (001). Phys. Rev. B. 47, 810 (1993)
51. Lin, T., Schwickert, M.M., Tomaz, M.A., Chen, H., Harp, G.R.: X-ray magnetic-circular-
dichroism study of Ni/Fe (001) multilayers. Phys. Rev. B. 59, 13911 (1999)
52. Aspelmeier, A., Tischer, M., Farle, M., Russo, M., Baberschke, K., Arvanitis, D.: Ac
susceptibility measurements of magnetic monolayers: MCXD, MOKE, and mutual induc-
tance. J. Magn. Magn. Mater. 146, 256 (1995).; Wu, J., Jin, X.F.: Temperature-dependent
magnetization in a ferromagnetic bilayer consisting of two materials with different Curie
temperatures. Phys. Rev. B 70, 212406 (2004)
53. Rader, O., Vescovo, E., Redinger, J., Blügel, S., Carbone, C., Eberhardt, W., Gudat, W.: Fe-
induced magnetization of Pd: the role of modified Pd surface states. Phys. Rev. Lett. 72, 2247
(1994)
54. Moruzzi, V.L.: Singular volume dependence of transition-metal magnetism. Phys. Rev. Lett.
57, 2211 (1986).; Moruzzi, V.L., Marcus, P.M., Schwarz, K., Mohn, P.: Ferromagnetic phases
of bcc and fcc Fe, Co, and Ni. Phys. Rev. B 34, 1784 (1986); Moruzzi, V.L., Marcus, P.M.:
Magnetism in bcc 3d transition metals: onset and approach to the Hund’s-rule limit. Phys.
Rev. B 38, 1613 (1988)
55. Guo, G.Y., Wang, H.H.: Gradient-corrected density functional calculation of elastic constants
of Fe, Co and Ni in bcc, fcc and hcp structures. Chin. J. Phys. 38, 949 (2000)
56. Tang, W.X., Qian, D., Wu, D., Wu, Y.Z., Dong, G.S., Jin, X.F., Chen, S.M., Jiang, X.M.,
Zhang, X.X., Zhang, Z.: Growth and magnetism of Ni films on GaAs (001). J. Magn. Magn.
Mater. 240, 404 (2002)
57. Tian, C.S., Qian, D., Wu, D., He, R.H., Wu, Y.Z., Tang, W.X., Yin, L.F., Shi, Y.S., Dong, G.S.,
Jin, X.F., Jiang, X.M., Liu, F.Q., Qian, H.J., Sun, K., Wang, L.M., Rossi, G., Qiu, Z.Q., Shi,
J.: Body-centered-cubic Ni and its magnetic properties. Phys. Rev. Lett. 94, 137210 (2005)
58. Schulz, B., Schwarzwald, R., Baberschke, K.: Magnetic properties of ultrathin Ni/Cu (100)
films determined by a UHV-FMR study. Surf. Sci. 307–309, 1102 (1994).; Zhang, R., Willis,
R.F.: Thickness-dependent Curie temperatures of ultrathin magnetic films: effect of the range
of spin-spin interactions. Phys. Rev. Lett. 86, 2665 (2001)
16 Metallic Magnetic Thin Films 843

59. Tischer, M., Arvanitis, D., Yokoyama, T., Lederer, T., Tröger, L., Baberschke, K.: Temperature
dependent MCXD measurements of thin Ni films on Cu (100). Surf. Sci. 307–309, 1096
(1994)
60. See Derek Craik: Magnetism. University of Nottingham, Nottingham (1995)
61. Thomas, O., Shen, Q., Schieffer, P., Tournerie, N., Lépine, B.: Interplay between anisotropic
strain relaxation and uniaxial interface magnetic anisotropy in epitaxial Fe films on (001)
GaAs. Phys. Rev. Lett. 90, 017205 (2003)
62. See, for example, Derek Craik: Magnetism: Principles and Applications, p. 392. Wiley,
Chichester (1995)
63. Mckeehan, L.W.: Magnetic interaction and resultant anisotropy in unstrained ferromagnetic
crystals. Phys. Rev. 52, 18 (1937)
64. Brooks, H.: Ferromagnetic anisotropy and the itinerant electron model. Phys. Rev. 58, 909
(1940)
65. Tarasov, L.P.: Ferromagnetic anisotropy of low nickel alloys of iron. Phys. Rev. 56, 1245
(1939)
66. Bozorth, R.M.: The permalloy problem. Rev. Mod. Phys. 25, 42 (1953)
67. James, P., Eriksson, O., Hjortstam, O., Johansson, B., Nordstrom, L.: Calculated trends of
the magnetostriction coefficient of 3d alloys from first principles. Appl. Phys. Lett. 76, 915
(2000)
68. Weinberger, P., Szunyogh, L., Blaas, C., Sommers, C., Entel, P.: Magnetic properties of bulk
Nic Fe1-c alloys, their free surfaces, and related spin-valve systems. Phys. Rev. B. 63, 094417
(2001)
69. Krebs, J.J., Jonker, B.T., Prinz, G.A.: Properties of Fe single-crystal films grown on (100)
GaAs by molecular-beam epitaxy. J. Appl. Phys. 61, 2596 (1987)
70. Yin, L.F., Wei, D.H., Lei, N., Zhou, L.H., Tian, C.S., Dong, G.S., Jin, X.F., Guo, L.P., Jia,
Q.J., Wu, R.Q.: Magnetocrystalline anisotropy in permalloy revisited. Phys. Rev. Lett. 97,
067203 (2006)
71. Noh, D.Y., Hwu, Y., Kim, H.K., Hong, M.: X-ray-scattering studies of the interfacial structure
of Au/GaAs. Phys. Rev. B. 51, 4441 (1995).; Andersson, T.G., Le Lay, G., Kanski, J.,
Svensson, S.P.: Room-temperature growth of two-dimensional gold films on GaAs (001).
Phys. Rev. B 36, 6231 (1987)
72. Wakelin, R.J., Yates, E.L.: A study of the order-disorder transformation in Iron-Nickel alloys
in the region FeNi3. Proc. Phys. Soc. London, Sect. B. 66, 221 (1953)
73. Wastlbauer, G., Bland, J.A.C.: Structural and magnetic properties of ultrathin epitaxial Fe
films on GaAs (001) and related semiconductor substrates. Adv. Phys. 54, 137 (2005)
74. Tian, C.S.: Ph. D. Thesis, Fudan University, Shanghai, China (2006)
75. Schumann, F.O., Willis, R.F., Goodman, K.G., Tobin, J.G.: Magnetic instability of ultrathin
fcc Fex Ni1-x films. Phys. Rev. Lett. 79, 5166 (1997)
76. Wimmer, E., Krakauer, H., Weinert, M., Freeman, A.J.: Full-potential self-consistent
linearized-augmented-plane-wave method for calculating the electronic structure of
molecules and surfaces: O2 molecule. Phys. Rev. B. 24, 864 (1981).; Weinert, M., Wimmer,
E., Freeman, A.J.: Total-energy all-electron density functional method for bulk solids and
surfaces. Phys. Rev. B 26, 4571 (1982)
77. Perdew, J.P., Burke, K., Ernzerhof, M.: Generalized gradient approximation made simple.
Phys. Rev. Lett. 77, 3865 (1996)
78. Shull, C.G., Wilkinson, M.K.: Neutron diffraction studies of the magnetic structure of alloys
of transition elements. Phys. Rev. 97, 304 (1955)
79. Wu, R.Q., Freeman, A.J.: Spin–orbit induced magnetic phenomena in bulk metals and their
surfaces and interfaces. J. Magn. Magn. Mater. 200, 498 (1999)
80. Dieny, B., Speriosu, V.S., Gurney, B.A., Parkin, S.S.P., Whilhoit, D.R., Roche, K.P., Metin,
S., Peterson, D.T., Nadimi, S.: Spin-valve effect in soft ferromagnetic sandwiches. J. Magn.
Magn. Mater. 93, 101 (1991)
81. Sands, T., Harbison, J.P., Leadbeater, M.L., Allen Jr., S., Hull, G.W., Ramesh, R., Keramidas,
V.G.: Epitaxial ferromagnetic τ-MnAl films on GaAs. Appl. Phys. Lett. 57, 2609 (1990)
844 D. Wu and X.-F. Jin

82. Tanaka, M., Harbison, J.P., Deboeck, J., Sands, T., Philips, B., Cheeks, T.L., Keramidas, V.G.:
Epitaxial growth of ferromagnetic ultrathin MnGa films with perpendicular magnetization on
GaAs. Appl. Phys. Lett. 62, 1565 (1993)
83. Matsui, M., Ido, T., Sato, K., Adachi, K.: Ferromagnetism and antiferromagnetism in Co–Mn
alloy. J. Phys. Soc. Jpn. 28, 791 (1970)
84. Rogers, D.J., Maeda, Y., Takei, K.: Compositional separation in Co-Mn magnetic thin films.
J. Appl. Phys. 78, 5842 (1995)
85. Thomson, T., Reidi, P.C., Wang, Q., Zabel, H.: 59 Co and 55 Mn NMR of CoMn alloys and
multilayers. J. Appl. Phys. 79, 6300 (1996)
86. Menshikov, A.Z., Takzei, G.A., Dorofeev, Y.A., Kazanstev, V.A., Kostyshin, A.K., Sych, I.I.:
Magnetic phase diagram of cobalt – manganese alloys. Zh. Eksp. Teor. Fiz. 89, 1269 (1985)
[Sov. Phys. JETP 62, 734 (1985)]
87. Wu, D., Liu, G.L., Jing, C., Wu, Y.Z., Loison, D., Dong, G.S., Jin, X.F., Wang, D.-S.:
Magnetic structure of Co1-x Mnx alloys. Phys. Rev. B. 63, 214403 (2001)
88. Jing, C., Wu, Y.Z., Yang, Z.X., Dong, G.S., Jin, X.F.: Structure and magnetism of Fe1−x Mnx
alloys on GaAs (001). J. Magn. Magn. Mater. 198–199, 270 (1999)
89. Hansen, M.: In: Hansen, M. (ed.) Constitution of Binary Alloys, p. 580. McGraw-Hill, New
York (1958)
90. Sumiyama, K., Yoshitabe, T., Nakamura, Y.: XPS valence band and core level spectra of
sputter-deposited Fe–Cu, Fe–Ag and Fe–Cu–Ag alloy films. J. Phys. Soc. Jpn. 58, 1725
(1989)
91. Chien, C.L., Liou, S.H., Kofalt, D., Wu, Y., Egami, T., McGuire, T.R.: Magnetic properties of
Fex Cu100-x solid solutions. Phys. Rev. B. 33, 3247 (1986)
92. Crespo, P., Hernando, A., Yavari, R., Drbohlav, O., García Escorial, A., Barandiarán, J.M.,
Orúe, I.: Magnetic behavior of metastable fcc Fe-Cu after thermal treatments. Phys. Rev. B.
48, 7134 (1993)
93. Ambrose, T., Gavrin, A., Chien, C.L.: Magnetic properties of metastable fcc Fe-Cu alloys
prepared by high energy ball milling. J. Magn. Magn. Mater. 124, 15 (1993)
94. Harris, V.G., Kemner, K.M., Das, B.N., Koon, N.C., Ehrlich, A.E., Kirkland, J.P., Woicik,
J.C., Crespo, P., Hernando, A., Garcia Escorial, A.: Near-neighbor mixing and bond dilation
in mechanically alloyed Cu-Fe. Phys. Rev. B. 54, 6929 (1996)
95. Manoharan, S.S., Klaua, M., Shen, J., Barthel, J., Jenniches, H., Kirschner, J.: Artificially
ordered Fe-Cu alloy superlattices on Cu (001). I. Studies on the structural and magnetic
properties. Phys. Rev. B. 58, 8549 (1998)
96. Uenishi, K., Kobayashi, K.F., Nasu, S., Hatano, H., Ishihara, K.N., Shingu, P.H.: Mechanical
alloying in the Fe-Cu system. Z. Metallkd. 83, 132 (1992)
97. Serena, P.A., García, N.: Ferromagnetism in FeCu metastable alloys. Phys. Rev. B. 50, 944
(1994)
98. Wang, J.-T., Zhou, L., Kawazoe, Y., Wang, D.-S.: Ab initio studies on the structural and
magnetic properties of FeCu superlattices. Phys. Rev. B. 60, 3025 (1999)
99. Mermin, M.D., Wagner, H.: Absence of ferromagnetism or antiferromagnetism in one-or two-
dimensional isotropic Heisenberg models. Phys. Rev. Lett. 17, 1133 (1966)
100. Pappas, D.P., Kamper, K.P., Hopster, H.: Reversible transition between perpendicular and
in-plane magnetization in ultrathin films. Phys. Rev. Lett. 64, 3179 (1990).; Pappas, D.P.,
Brundle, C.R., Hopster, H.: Reduction of macroscopic moment in ultrathin Fe films as the
magnetic orientation changes. Phys. Rev. B 45, 8169 (1992)
101. Qiu, Z.Q., Pearson, J., Bader, S.D.: Asymmetry of the spin reorientation transition in ultrathin
Fe films and wedges grown on Ag (100). Phys. Rev. Lett. 70, 1006 (1993)
102. Wu, Y.Z., Won, C., Scholl, A., Doran, A., Zhao, H.W., Jin, X.F., Qiu, Z.Q.: Magnetic stripe
domains in coupled magnetic sandwiches. Phys. Rev. Lett. 93, 117205 (2004).; Won, C.,
Wu, Y. Z., Choi, J., Kim, W., Scholl, A., Doran, A., Owens, T., Wu, J., Jin, X.F., Qiu, Z.Q.:
Magnetic stripe melting at the spin reorientation transition in Fe/Ni/Cu (001). Phys. Rev. B
71, 224429 (2005)
16 Metallic Magnetic Thin Films 845

103. Schulz, B., Baberschke, K.: Crossover from in-plane to perpendicular magnetization in
ultrathin Ni/Cu (001) films. Phys. Rev. B. 50, 13467 (1994)
104. Farle, M., Mirwald-Schulz, B., Anisimov, A.N., Platow, W., Baberschke, K.: Higher-order
magnetic anisotropies and the nature of the spin-reorientation transition in face-centered-
tetragonal Ni (001)/Cu (001). Phys. Rev. B. 55, 3708 (1997)
105. Matthes, F., Seider, M., Schneider, C.M.: Strain-induced magnetic anisotropies in ultrathin
epitaxial Nix Pd1-x alloy films. J. Appl. Phys. 91, 8144 (2002)
106. Yu, P., Yin, L.F., Wei, D.H., Tian, C.S., Dong, G.S., Jin, X.F.: Correlation between spin
reorientation transition and Curie temperature of Nix Pd1-x alloy on Cu (001). Phys. Rev. B.
79, 212407 (2009)
107. Oepen, H.-P., Knappmann, S., Wulfhekel, W.: Ferro-and para-magnetic properties of ultrathin
epitaxial Co/Cu films. ibid. 148, 90 (1995).; Garreau, G., Farle, M., Beaurepaire, E.,
Baberschke, K.: Curie temperature and morphology in ultrathin Co/W (110) films. Phys. Rev.
B 55, 330 (1997)
108. Arnold, C.S., Venus, D.: Simple window-compensation method for improving the signal-to-
noise ratio in measurements of the magneto-optic Kerr effect in ultrathin films. Rev. Sci.
Instrum. 66, 3280 (1995)
109. Arnold, C.S., Johnston, H.L., Venus, D.: Magnetic susceptibility measurements near the
multicritical point of the spin-reorientation transition in ultrathin fcc Fe (111)/2 ML Ni/W
(110) films. Phys. Rev. B. 56, 8169 (1997)
110. O’Brien, W.L., Tonner, B.P.: Transition to the perpendicular easy axis of magnetization in Ni
ultrathin films found by x-ray magnetic circular dichroism. Phys. Rev. B. 49, 15370 (1994)
111. Farle, M., Platow, W., Anisimov, A.N., Poulopoulos, P., Baberschke, K.: Anomalous reorien-
tation phase transition of the magnetization in fct Ni/Cu (001). Phys. Rev. B. 56, 5100 (1997)
112. Fritzsche, H., Kohlhepp, J., Elmers, H.J., Gradmann, U.: Angular dependence of perpendicu-
lar magnetic surface anisotropy and the spin-reorientation transition. Phys. Rev. B. 49, 15665
(1994)
113. Heinrich, B., Monchesky, T., Urban, R.: Role of interfaces in higher order angular terms of
magnetic anisotropies: ultrathin film structures. J. Magn. Magn. Mater. 236, 339 (2001)

Di Wu is a professor in Department of Physics, Nanjing Uni-


versity, China. He received his B.S. (1997) and PhD (2001) at
Fudan University, China. He had 6 years postdoctoral research
experience in University of Utah and University of California,
Riverside. His recent research focuses on spin current transport
in magnetic insulators.
846 D. Wu and X.-F. Jin

Xiaofeng Jin is a Professor in Department of Physics, Fudan


University, China. He received his PhD at Fudan University in
1989. He served as the chair of the International Colloquium
on Magnetic Films and Surfaces in 2012 and the International
Union of Pure and Applied Physics Magnetism Commission C9
in 2014–2017. His research focuses on spin transport in thin film.
Magnetic Oxides and Other Compounds
17
J. M. D. Coey

Contents
Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 848
Rocks, Solid Solutions, and Percolation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 851
Principles of Oxide Magnetism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 852
Iron Oxides and Hydroxides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 863
Hematite . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 864
Magnetite . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 866
Maghemite . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 867
Wüstite . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 867
Iron Hydroxides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 868
Ferrites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 869
Spinels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 869
Garnets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 871
Orthoferrites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 872
Hexagonal Ferrites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 873
Other Magnetic Oxides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 875
3d Oxides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 876
Perovskites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 878
Pyrochlores . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 885
4d and 5d Oxides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 887
4f Oxides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 888
5f Oxides and Related Compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 889
Related Compounds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 890
Halides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 890
Chalcogenides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 892
Pnictides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 895
Silicates and Carbonates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 897
Silicates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 897
Carbonates, Phosphates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 903

J. M. D. Coey ()
School of Physics, Trinity College, Dublin, Ireland
e-mail: jcoey@tcd.ie

© Springer Nature Switzerland AG 2021 847


J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_17
848 J. M. D. Coey

Oxide Thin Films . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 903


Substrates, Caps, and Buffers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 903
Thin Film Preparation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 904
Magnetic Oxide Monolayers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 909
Oxide Heterostructures and Interfaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 914
Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 917
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 917

Abstract

Magnetic oxides are important functional magnetic materials, both as permanent


magnets and as soft magnetic materials for high-frequency and microwave appli-
cations. They have been central in developing our understanding of exchange in
strongly correlated insulating and barely metallic systems, including the effects
of magnetic frustration and disorder. Oxide thin films are playing an increasing
role in spin electronics. This chapter begins with an outline of the natural
distribution of magnetic elements, with an initial emphasis on naturally occurring
oxide minerals, especially oxides, hydroxides, and silicates of iron and other
magnetic cations. Principles of 3d oxide magnetism and crystal chemistry are
introduced. Structure and intrinsic magnetic properties of the main structural
families of binary and ternary oxides and related compounds, including halides,
chalcogenides, and pnictides, are then presented in tabular form. Magnetism of
4d, 5d, and 5f oxides is also covered. Then methods of preparing oxide thin
films are presented, and specific features of single oxide films, interfaces, and
heterostructures are discussed.

Background

Terrestrial minerals are the source of the magnetic elements needed to make all
the alloys and compounds of interest in this handbook. The ores are culled from the
Earth’s surface or the topmost kilometer of the crust, where low-density compounds
of oxygen and other electronegative elements segregate, while denser siderophilic
materials tend to sink deep into the core, which is composed of 90% iron with a
little nickel – a composition similar to that of Ni-Fe meteorites. In fact, iron is the
most abundant element by weight in the whole Earth (35 wt%), but in the crust
it occupies fourth position (6 wt%), after oxygen, silicon, and aluminum, or sixth
position, after hydrogen and sodium if its concentration is expressed in atom percent
(2.1 at%). Crustal abundances in at% are shown in Fig. 1a). Rock-forming minerals
are therefore predominantly aluminosilicates, but two iron oxides, magnetite and
hematite, feature among them. These were the first magnetic materials known to
man, and they remain our primary source of iron, which provided the basis of
technology and weaponry for over 2000 years.
17 Magnetic Oxides and Other Compounds 849

a)

5 Fe
Log (abundance, ppm)

4
Mn
3
Cr Ni Ce
2 Nd
Co
Pr Sm Gd
1 Dy
Er Yb
Eu Ho
Tb
0 Pm Tm

-1
3d elements 4f elements
b)

Fig. 1 (a) Atomic abundances of the principal elements in the Earth’s crust and (b) abundances
of magnetic 3d and 4f elements, plotted on a logarithmic scale

Almost 1 atom in 40 in the crust is iron, and compared with other magnetic
elements, the crustal abundance of iron is 40 times that of all the others put together.
This 40/40 rule affords some perspective on the magnetism of natural oxides.
Abundances of the magnetic elements are compared on a log scale in Fig. 1b), where
it can be seen that there is about three orders of magnitude more Fe in the Earth’s
crust than either of the other ferromagnetic 3d elements Co and Ni (Table 1).
The magnetic elements, which have ions with unpaired electrons, are found
predominantly among the 3d and 4f series. The light rare earths are not particularly
rare – neodymium is as abundant as cobalt or zinc, but the abundances follow
an odd/even variation with atomic number that reflects the stability of the atomic
nuclei, superposed on a decline with increasing atomic number. The heavier 4f
elements become increasingly uncommon. Crustal abundance is a rough proxy
850 J. M. D. Coey

Table 1 Elements in the Ion Abundance Configuration


Earth’s crust (atomic %)
O2− 59.7 2p6
Si4+ 20.9 2p6
Al3+ 6.1 2p6
H+ 2.9 –
Na+ 2.2 2p6
Fe2+/3+ 2.1 3d6/5
Ca2+ 2.1 3p6
Mg2+ 2.0 2p6
K+ 1.1 3p6

1.0 o o o o o o o o o o o o

o o o o o o o o o o o o
0.8 o o o o o o o o o o

o o o o o o o o o o o
Tc/Tc (x= 1)

0.6 o o o o o o o o o
Cluster
o o o o o o o o o o
0.4 o o o o o o o o o o o o

o o o o o o o o o o
0.2 xp Antiferro-
o
magnetico o o o o o o o
pair Isolated
o o o o o o o o o o o o
0.0 o o o o o o o o o o o
0.0 0.2 0.4 0.6 0.8 1.0
x

Fig. 2 (a) Schematic variation of the magnetic ordering temperature as a function of the
concentration x of magnetic ions in an oxide solid solution of magnetic and nonmagnetic cations
M and N, (Mx N1-x )Oη , where η is a small integer or rational fraction. The magnetic interactions
couple nearest-neighbor magnetic cations, and xp is the percolation threshold. The distribution of
ions illustrated in (b) well below the percolation threshold

for cost. The periodic table in Table 2 provides an indication of the costs of the
elements in five broad bands. It is truly serendipitous for magnetism that the most
abundant and cheapest magnetic element turns out to be the one that is ferromagnetic
with the highest magnetization and the one that forms 3d cations with the greatest
magnetic moments. The natural magnetic materials that feature in areas such as rock
magnetism or biomineralization are iron oxides.
Oxides exhibit a wide range of electric, magnetic, optical, and structural proper-
ties. They can be insulating, semiconducting, metallic, ferroelectric, piezoelectric,
ferromagnetic, half-metallic, ferrimagnetic, antiferromagnetic, or superconducting.
From a magnetic point of view, the dominant interaction in transition-metal oxides is
usually antiferromagnetic superexchange, the coupling of the spins of two 3d cations
separated by an oxygen anion, which leads to antiferromagnetic or ferrimagnetic
order. Ferromagnetism is less common in oxides, where it is often associated with
cation mixed-valence and orbital degeneracy, than it is in metals. See  Chap. 2,
“Magnetic Exchange Interactions.”
17 Magnetic Oxides and Other Compounds 851

Table 2 Periodic table of the elements, showing their approximate costs per kilogram, after [19]

Rocks, Solid Solutions, and Percolation

Oxide structures have great scope for accommodating cations of similar charge
and ionic radius, and their properties can be considerably modified thereby. In
some cases, there is an extended range or even complete solid solubility between
chemically and structurally similar end-members.
Pure, single-phase end-members are not the norm in nature. The relative abun-
dance of Al and Fe in the crust (Fig. 1) means that iron (and other magnetic cations)
are relatively sparse in aluminosilicate rocks, which are generally multiphase
aggregates that reflect the temperature, pressure, and chemical environment in which
the rock was formed. It may be possible to separate out one of the phases for
examination, by physical or magnetic means. Otherwise, the whole multiphase rock
or soil can be analyzed microscopically in order to characterize the natural mixture
of minerals present.
Iron is often an impurity in aluminosilicate minerals, replacing aluminum
if trivalent or an alkaline earth cation if divalent. Iron-rich end-members are
the exception. When the iron content in a solid solution is low, superexchange
interactions among ferrous or ferric cations separated by a single oxygen anion
are inhibited, and long-range magnetic order is impossible. The magnetically dilute
minerals then exhibit Curie-Weiss paramagnetism. Percolation becomes possible
when the iron concentration x on appropriate sites exceeds the percolation threshold
852 J. M. D. Coey

xp , which depends on the crystal structure. A rough estimate of xp is 2/Z, where


Z is the nearest-neighbor cation coordination number – 8 if the cations form a bcc
lattice, 12 for an fcc lattice, 4 for a square lattice, etc. Magnetic order appears when
x > xp , and the transition temperature increases with x, although the transition itself
is broadened by the disorder (Fig. 2).
Natural iron-rich, end-member oxides tend to be contaminated by aluminum,
manganese, titanium, chromium, or other ions, which can modify the magnetic
properties significantly; common examples are Al in hematite and Ti in magnetite.
When discussing natural oxides here, we will often cite ideal chemical formu-
lae, which differ from the more complex chemical composition of real natural
samples. But many of these minerals are of interest in their own right; naturally
occurring magnetic oxides have found important applications in paleomagnetism or
archaeomagnetism, for example, and there are interesting natural examples of low-
dimensional or mixed-valence magnetic materials. Nature, with its access to high
pressures and geological timescales, provides us with a rich store of natural crystals
that we are unable to replicate in the laboratory. Tetrataenite, the tetragonal, ordered
L10 -structure FeNi alloy found in nickel-iron meteorites that have equilibriatesd
over a huge timespan, is an example.
Phase purity should not be a problem in synthetic oxides, but deliberate cation
substitution can be a powerful tool to modify exchange interactions, conductivity,
and sublattice magnetization in order to achieve optimal magnetic properties or to
investigate how magnetism depends on composition.

Principles of Oxide Magnetism

Oxides are ionic compounds where small, highly charged metal cations are embed-
ded in a lattice of larger oxygen anions that occupy 90% of the volume. Bonding is
predominantly ionic. Oxygen accepts two electrons to form the divalent O2− anion
with the particularly stable 2p6 closed shell, a configuration that is shared by 90% of
all the ions in the Earth’s crust. The extra electrons are transferred to oxygen from
the metals that form the positively charged cations; those listed in Table 2 are the
ones most commonly encountered in oxides and other magnetic compounds. These
cations are all smaller than the oxygen anion, for which the ionic radius is taken to be
140 pm. Other anions encountered in magnetic compounds include OH− (110 pm),
F− (133 pm), Cl− (181 pm), S2− (184 pm), and N3− (146 pm). The ionic picture is,
of course, oversimplified. The chemical bond between metal and oxygen has part-
ionic and part-covalent character, governed by the electronegativity of the metal
partner. Covalency is more pronounced in sites with a low coordination number and
for cations like Si4+ or V4+ with a high formal charge state, as well as for anions
such as S2− and N3− that are much less electronegative than oxygen or fluorine
(Table 3).
The crystal structures of most oxides may be regarded as a packing of a lattice of
large oxygen anions held together by electrostatic interactions with the small metal
cations in interstitial sites. The structures sometimes incorporate large cations that
17 Magnetic Oxides and Other Compounds 853

Table 3 Ionic radii of cations in oxides. Values in brackets refer to a low-spin state
Fourfold Sixfold Sixfold Eightfold
tetrahedral pm octahedral pm octahedral pm cubic pm
Mg2+ 57 V4+ 3d1 58 Ti3+ 3d1 67 Ca2+ 112
Zn2+ 60 Cr4+ 3d2 55 V3+ 3d2 64 Sr2+ 126
Al3+ 39 Mn4+ 3d3 53 Cr3+ 3d3 62 Ba2+ 142
Fe3+ 3d5 49 Mn2+ 3d5 83 Mn3+ 3d4 65 Pb2+ 129
Si4+ 26 Fe2+ 3d6 78 (61) Fe3+ 3d5 65 Y3+ 102
Co2+ 3d7 75 (65) Co3+ 3d6 61 (55) La3+ 116
Ni2+ 3d8 69 (56) Ni3+ 3d7 60 (56) Gd3+ 4f7 105

substitute on oxygen sites. The packing may be body-centered cubic (perovskites),


face-centered cubic (spinels, garnets), or hexagonal close-packed (hexaferrites,
corundums). Interstices in these packings are normally of two types – tetrahedral
and octahedral – which explains the emphasis on these two sites permeates any
discussion of oxide magnetism. They are illustrated for the ideal face-centered cubic
structure in Fig. 3, where both interstitial sites in the close-packed lattice have cubic
symmetry. The radius of an ion that will just fit into a close-packed tetrahedral

interstice is rtet = ( (3/2)−1)rO = 29 pm where rO = 140 pm is the ionic radius
of O2− . The close-packed octahedral interstice is larger; it can just accommodate an

ion of radius roct = ( 2 – 1)rO = 52 pm. Nevertheless, the magnetic cations listed
in Table 2 are too big to fit into tetrahedral sites and even a bit too big to snugly
occupy octahedral sites without distorting the oxygen lattice. The distortion may be
uniform, retaining the cubic site symmetry, or uniaxial along a threefold or fourfold
axis of the tetrahedron or octahedron.
The 4f cations are larger, and they are eight- to twelve- fold coordinated in
oxides. Rare-earth cations are usually trivalent, and their ionic radii in eightfold
cubic coordination decrease from 116 pm for La3+ to 98 pm for Lu3+ . Eu2+ and
Yb2+ are exceptions; they have a stable half-filled (4f7 ) or filled (4f14 ) 4f shells,
with radii of 125 pm and 114 pm, respectively.
The electrostatic coulomb interaction among the 3d electrons of a magnetic
cation and between them and the negatively charged 2p6 shells of their oxygen anion
neighbors are the fundamental interactions in oxide magnetism. A critical first step
is to replace the spherically symmetric d wave function of the free ions with linear
combinations that reflect the local crystal symmetry. A drastic simplification is the
crystal field model, where we replace the oxygen anions by -2e point charges at the
anion sites and consider the energy of the electrons in the unfilled 3d shell to see
how the crystal field created by the O2− neighbors influences both the electronic
structure and magnetic properties. The one-electron model [13], which ignores the
on-site 3d-3d Coulomb interactions, is a fair approximation for d1 , d4 , d6 , and d9
ions when Hund’s first rule applies because these ions have a single electron or hole
outside an empty (d0 ), half-filled (d5 ), or filled (d10 ) d-shell.
854 J. M. D. Coey

Fig. 3 Tetrahedral and


octahedral sites in a
cubic-close-packed oxygen
lattice

When the tetrahedral or octahedral site is undistorted, the free ion eigenfunctions
ψ0 , ψ±1 , and ψ±2 , with subscripts denoting the orbital magnetic quantum number
ml , are replaced by suitable linear combinations that reflect the cubic site symmetry.
They are:

 √ 
ψxy = −i/ 2 (ψ2 − ψ−2 )
 √ 
ψyz = i/ 2 (ψ1 + ψ−1 )
 √ 
ψzx = −1/ 2 (ψ1 − ψ−1 ) (1)
 √ 
ψx2−y2 = 1/ 2 (ψ2 + ψ−2 )
ψz2 = ψ0

Figure 4 compares the electron density for this basis set of d orbitals in an
octahedral environment with that of the d electrons in a spherical potential.
On inspecting the disposition of the oxygen electrons in the surrounding
octahedron, it is obvious that the xy, yz, and zx orbitals are degenerate – they
are labeled t2g orbitals. It is less obvious that the x2 – y2 and z2 orbitals, labeled
eg , are degenerate, but they are clearly higher in energy because the 3d electron
density is maximum near the negatively charged anions. The crystal field splitting
in octahedral coordination (where the suffix g is included to indicate the presence of
a center of symmetry) is illustrated in Fig. 5.
17 Magnetic Oxides and Other Compounds 855

Fig. 4 The 3d orbitals for electrons in a free ion (top row) and the orbitals for the ion in a cubic
crystal field (bottom row). Oxygen positions in a surrounding octahedron are marked by red dots
that represent -2e point charges

Fig. 5 Splitting of the t2g


and eg orbitals in an
octahedral environment. cf
is the crystal field splitting
calculated in a point charge
model. The extra ligand field
splitting due to the
hybridization of 3d and sp
orbitals gives the total
splitting 

The splitting of the one-electron levels is only partly explained by the electro-
static potential created by the oxygen anions; about half of it should be attributed
to covalency associated with the different overlap of the t2g and eg orbitals with
the lower-lying 2p oxygen orbitals, which introduces a greater bonding-antibonding
splitting for the 2p-eg σ-bonds than the 2p-t2g π-bonds. This is the ligand field effect.
The combined ionic and covalent effects result in a total splitting , which is about
1 eV in octahedral sites [5].
The crystal field splitting is reversed in pure cubic coordination, where the eg
orbitals are lowest. The splitting is smaller and also reversed in tetrahedral sites,
with the e orbitals lower and the t2 orbitals higher. The suffix g is dropped because
a tetrahedron has no center of symmetry (Fig. 6).
The influence of a distortion of the cubic site is to lift the degeneracy of the
one-electron energy levels, as shown in Fig. 7. The splitting preserves the center of
gravity of each set of orbitals.
Whenever a set of orbitals is not fully occupied, the crystal field may lower the
energy of the ion by an amount known as the crystal field stabilization energy cfse ,
which can dictate the site preference of ions in structures where there is a choice of
856 J. M. D. Coey

Fig. 6 Energy splitting of the one-electron states in cation sites with octahedral, tetrahedral, and
cubic coordination due to the crystalline electric field

octahedral or tetrahedral site occupancy. For example, the energy of Cr3+ (3d3 ) in
octahedral sites is 3 × (−2/5) O = −1.2 O , whereas for Fe3+ (3d5 ) it is zero.
Cr3+ therefore has a strong octahedral site preference. Furthermore, a distortion of
the environment can provide an additional stabilization. These considerations can
outweigh purely steric effects.
The Jahn-Teller effect is the tendency of some ions to spontaneously distort their
local environment in order to gain some extra crystal field stabilization energy. A 3d3
ion in an octahedral site gains nothing from a distortion of the oxygen octahedron
because of the center of gravity rule, but a 3d4 ion will tend to induce a local
tetragonal deformation which splits the eg levels, lowering the energy of either the
occupied dz2 or dx2–y2 orbital (Fig. 7). The Jahn-Teller effect is strongest for d4 and
d9 ions in octahedral coordination and d1 and d6 ions in tetrahedral coordination.
Two other important electronic effects encountered in 3d oxides are charge order
and orbital order [62]. Here interionic coulomb interactions play a role. Charge
17 Magnetic Oxides and Other Compounds 857

z z

x x x

x x x

z z z

αt

Δ cfse

Δ0

δt

Fig. 7 Influence of a tetragonal elongation or flattening of an octahedral site on the splitting of


the one-electron energy levels. The crystal field stabilization energy cfse is indicated for a 3d4 ion
such as Mn3+ with a t2g 3 eg (x2 − y2 ) configuration

order arises when a particular lattice site can be occupied by an ion in one of
two different charge states. Examples are Mn3+ and Mn4+ or Fe2+ and Fe3+ . The
average 3d occupancy is nonintegral, but the electrons may settle in an ordered
pattern on the cation sites in a charge-ordered state with some lattice distortion.
Otherwise, the surplus 3d electrons rapidly hop or tunnel among the d3 or d5
ion cores, a phenomenon known as mixed valence that occurs, for example, in
mixed-valence manganites such as (La0.7 Sr0.3 ){Mn3+ 0.7 Mn4+ 0.3 }O3 or magnetite
[Fe3+ ]{Fe2+ ,Fe3+ }O4 , where the brackets denote different crystallographic sites.
The average 3d occupancy on octahedral sites {} in these examples is 3.7 or 5.5,
respectively.
Orbital order arises when the 3d occupancy is integral and single-valued, but the
electron occupies a degenerate orbital. An example is Mn3+ (t2g 3 eg 1 ). A lattice
distortion such as tetragonal compression and expansion on alternate sites can lead
to alternate occupancy of dz2 and dx2-y2 orbitals, in a variant of the Jahn-Teller
effect. The ordered electronic state generally reverts to a disordered state at a high-
temperature phase transition, where the entropy of disorder Rln2 per mole that is
released is comparable to the entropy Rln(2S + 1) released in the vicinity of a
magnetic order-disorder phase transition.
858 J. M. D. Coey

Coulomb interactions among electrons in the single electron states are critical,
both within and between cations. Although the one-electron picture is rather easy
to grasp and relate to energy-band calculations, it does no justice to these many-
electron interactions in 3d ions, especially for high-spin 3d ions with 2, 3, 7, or 8
electrons, which are in an F-state. Ions with five electrons are in an S state, which
corresponds to a half-filled 3d shell with spherical symmetry, while those with one,
four, six, or nine electrons are in an D-state, which maps onto a t2 or e one-electron
state. Here S, D, and F refer to the values 0, 2, or 3 for the total orbital angular
momentum L of the free ion allowed by Hund’s second rule, discussed below. The
excited states of the 3dn shell which are probed in optical transitions have been
calculated from crystal field theory, and they are represented on Tanabe-Sugano
diagrams [101, 108]. In magnetism, it is usually sufficient to consider only the
ground state.
The intra-atomic interactions give rise to Hund’s rules for the spin and orbital
moments of the many-electron ions. The first rule gives the net spin S, which is the
maximum value, consistent with the Pauli principle that no two electrons can occupy
the same quantum state with the same spin. It is simply obtained by populating
the lowest orbitals with the available electrons. The interaction is represented by
exchange coupling JH of a pair of electrons in different orbitals with the same spin,
written in terms of an atomic Hamiltonian

 
1
HH = −JH α=β + 2s iα .s iβ (2)
2

where typical values of the intra-atomic exchange JH are 0.8–0.9 eV for 3d electrons
and 0.6–0.7 eV for 4d electrons. Values for the more compact 4f shell are higher.
Here, s is the spin of an electron in a particular orbital at site i and the sum is over all
electron pairs on the site. Two examples, for d4 and d6 , are shown in Fig. 8a and b.
The energy contribution to (2) is JH for a ↑↑ pair and zero for a ↑↓ pair, so we have
-6JH for d4 and -10JH for d6 . Hund’s first rule implies that the five ↓ orbitals lie
above the five ↑ orbitals, which is normally the case unless the crystal/ligand field
splitting is large or the on-site exchange is weak when it is possible to have a low-
spin state, where in octahedral coordination, for example, the t2g ↓ orbitals are filled
before the eg ↑ orbitals and Hund’s first rule no longer applies. This is illustrated
in Fig. 8c and d, where the total spin moments are, respectively, S = 1 and S = 0.
Low-spin states in oxides are sometimes found for Co3+ in octahedral coordination,
which has S = 2 (t2g 3↑ eg 2↑ t2g 1↓ ) in the high-spin state and S = 0 (t2g 3↑ t2g 3↓ ) in
the low-spin state. They are more common in more covalent compounds, where the
ligand field is larger.
Hund’s second rule maximizes the orbital angular momentum consistent with the
value of S for the ion. The possibilities in Table 4 are a D-state with L = 2(d1 , d4 ,
d6 , d9 ), an F-state with L = 3(d2 , d3 , d7 , d8 ), or an S-state with L = 0 (d5 ). The
17 Magnetic Oxides and Other Compounds 859

Fig. 8 Electronic configurations of an ion in octahedral coordination having four or six electrons.
In (a) and (b), the usual high-spin states are illustrated, whereas (c) and (d) show the low-spin
states that arise when the crystal field splitting cf exceeds the intra-ionic Hund’s rule exchange
splitting H

Table 4 Ground state terms for 3d and 4f ions, 2S + 1 LJ deduced from Hund’s rules
3d1 3d2 3d3 3d4 3d5 3d6 3d7 3d8 3d9
2D 3F 4F 5D 6S 5D 4F 3F 2D
3/2 2 3/2 0 5/2 4 9/2 4 5/2

4f1 4f2 4f3 4f4 4f5 4f6 4f7 4f8 4f9 4f10 4f12 4f12 4f13
2F 3H 4I 5I 6H 7F 8S 7F 6H 5 4I 3H 2F
5/2 4 9/2 4 5/2 0 7/2 6 15/2 I8 15/2 6 7/2

nomenclature here is based on free ions, but it is retained for ions in solids, even
though the orbital moment is almost completely quenched by the crystal field.
The third rule invokes spin-orbit coupling, a relativistic interaction of great
interest in magnetism, represented by the Hamiltonian:

Hso = L.S (3)

where  is the ionic spin-orbit coupling constant. In 3d ions with quenched


orbital moments, the effect of the spin-orbit interaction is to restore a small orbital
contribution, thereby creating magnetocrystalline anisotropy. Octahedral Co3+ is a
good example.
The crystal field in 4f ions is small in comparison with the spin-orbit interaction,
so the orbital moment is unquenched and L and S couple together to form the total
angular momentum J = L ± S, just as they do in free ions.
We now consider briefly the intra-atomic coulomb correlations. Unlike the delo-
calized, nearly free electrons in broad energy bands in semiconductors and many
metals, the 3d electrons in oxides are usually localized, with an integral number
860 J. M. D. Coey

4s(T)
conduction band
3dn±1
U W
3dn

3dn+1+ L
U W
2p6(O) 3dn
valence band

Fig. 9 Schematic density of states for a 3d oxide, showing the charge-transfer process that leads
to a Mott insulator in an early transition-metal oxide and a charge-transfer insulator in a late
transition-metal oxide. The oxide is insulating if the excited state cannot be accommodated within
the bandwidth

of them associated with each cation site. The 3d levels, illustrated schematically
in Fig. 9, lie in the gap between a broad, full 2p(O) valence band and a broad,
empty 4s(T) conduction band, where T is the 3d transition metal. The 3d levels
lie near the top of the bandgap for cations like Ti3+ or V3+ at the beginning of
the series and move toward the bottom of the gap for cations like Ni2+ or Cu2+ at
the end. The 3d orbitals inevitably overlap to some extent, and in the tight-binding
approximation with interatomic hopping integral t, this means that they form a
narrow band of width W = 2Zt, where Z is the coordination number. The bandwidth
decreases as the nuclear charge increases across the 3d series. Elementary band
theory unphysically predicts that any partially filled band, however narrow, will
always be conducting, because the theory neglects electronic correlations. Oxides
are not usually metals, despite a 3d bandwidth of 2 eV or more, for reasons
explained by Mott [75]. In a metal, the average 3dn configuration has to coexist with
fluctuating 3dn ± 1 configurations in order for conduction to take place. The energy
cost of transferring an electron from one site to the next is the coulomb energy
Udd , which is the difference between the ionization energy and the electron affinity
of a 3dn ion. When this exceeds W, Udd /W > 1, conduction cannot occur, and the
oxide is known as a Mott insulator (or a Mott-Hubbard insulator). The d-d coulomb
correlations transform what would otherwise be a metal into an insulator, because
there is no place in the band to accommodate the correlation energy.
For the heavier 3d ions, an easier charge-transfer process is from a filled oxygen
2p6 shell to an adjacent metal cation, creating a 2p5 configuration with a ligand
hole and a 3dn + 1 cation configuration with an excess electron. This process costs
an energy Upd , and when Udd > Upd > W, the oxide is known as a charge-transfer
insulator.
17 Magnetic Oxides and Other Compounds 861

The physics of Mott insulators is captured by the Hubbard’s deceptively simple-


looking Hamiltonian for an s band. It is a favorite toy of condensed-matter theorists.

H = −t<ij >,α c† iσ cj σ + U j ni↑ ni↓ (4)

Here the first term transfers an electron from site i to a neighboring site j, without
flipping the spin. The second term is the on-site coulomb repulsion between two
electrons of opposite spin on the same site. The nearest-neighbor interaction is
antiferromagnetic because the wave function cannot spread out to neighboring sites,
unless the spins are antiparallel. The coupling energy between the spins on adjacent
sites is of order t2 /U.
Modern electronic structure calculations, especially those based on density
functional theory, have been good at predicting the magnetization and magnetic
ordering temperature of metals, but they are less successful with oxides and other
insulators, where the coulomb interaction parameter U is often has to be chosen to
match the observed bandgap. Numerical calculations dispense with the simplified
models that have helped us to develop an intuitive physical understanding of the
many-electron ground state in terms of a handful of physical parameters that
quantify the exchange, spin-orbit, and crystal field interactions that really call the
shots in this area of magnetism.
Oxides are rarely perfectly stoichiometric, so exactly integral d-band occupancy
is unlikely. There will inevitably be spatial charge fluctuations associated with
atomic defects, but they do not normally spoil the insulating state. Local lattice
deformations arise because the 3d electrons or holes interact strongly with their
ionic environment and create lattice polarons with a large effective mass that are
effectively immobile. Furthermore, if an electron should hop to an unoccupied
neighboring site in a 2D or 3D antiferromagnetic array, it leaves a hole behind
and flips a spin with each subsequent hop, creating a ferromagnetic interaction
while leaving a trail of magnetic defects behind it. The energy can be mini-
mized by the electron returning to its starting point. A small excess of electrons
or holes in the half-filled Hubbard band does not therefore make the model
conducting.
Direct overlap of the 3d orbitals of cations on neighboring sites to form the
3d band is relatively small, but the mixing of 3d states with the 2p states of
neighboring oxygen anions is more significant. In these circumstances, the exchange
interactions coupling the spins of next-neighbor cations are indirect, superexchange
interactions. These were systematically investigated by Goodenough and Kanamori
[50] in the 1960s. Their rules were simplified by Anderson [1] and the interactions
are discussed in  Chap. 2, “Magnetic Exchange Interactions.” The coupling is
usually antiferromagnetic, except if the M-O-M bond angle is close to 90◦ , when
weak ferromagnetic interactions can occur. The antiferromagnetic interactions lead
to either antiferromagnetic or ferrimagnetic order, depending on the lattice structure,
but they may be frustrated by the lattice topology.
862 J. M. D. Coey

In mixed-valence compounds, where the 3d electrons are delocalized, the elec-


tron hopping with spin memory leads to the ferromagnetic double exchange inter-
action, first described by Zener [116]. Both superexchange and double exchange are
described by the Heisenberg Hamiltonian:

H = −2JS i .S j (5)

where a positive or negative sign of the exchange parameter J corresponds to a


ferromagnetic or an antiferromagnetic intersite exchange interaction. Si is the total
spin localized on site i.
Exchange interactions with different symmetry are usually much weaker. The
Dzyaloshinskii-Moriya (D-M) interaction, for example, which has the form

 
HDM = −D. S i × S j (6)

where the vector D must lie along a uniaxial symmetry axis, was discovered in
response to the problem of canted antiferromagnetism in hematite [74] and some
other antiferromagnetic insulators.
Finally, Fig. 10 provides an overview of the magnetic ordering temperatures in
oxides. The histogram illustrates the distribution of Curie and Néel temperatures for

Fig. 10 Magnetic transition temperatures for oxides. Ferromagnets are shown in blue, ferrimag-
nets in pink, and antiferromagnets in red. Data are based on [33]
17 Magnetic Oxides and Other Compounds 863

a thousand magnetically ordered oxides. None of them exceeds 1000 K. Oxides of


ions Fe3+ or Mn2+ with a 3d5 configuration and the maximum value S = 5/2 tend
to order at the highest temperatures. Exchange between rare-earth ions in oxides is
weak, on account of the limited spatial extent of the 4f wave functions and the large
value of the on-site coulomb interaction U ∼ 6 eV. Hence t2 /U is small.
The next five sections present the different families of magnetic oxides and
related compounds. We use the structure types of natural minerals as a frame of
reference to discuss a wide range of synthetic materials that are never found in the
natural environment.

Iron Oxides and Hydroxides

Iron oxides and hydroxides constitute the most important group of magnetic
minerals [35]. Most important are the famous magnetic rock-forming oxides
magnetite and hematite. Then there is a series of crystalline or partly amorphous iron
oxyhydroxides and hydroxides that are found in soils and sediments. A summary of
the structural and intrinsic magnetic properties of the iron end-members of whole
group is given in Table 5; substitution of other magnetic or nonmagnetic cations
(especially Al3+ ) in the iron end-members will modify these values. Variability is
ubiquitous in natural specimens; no two examples of a mineral with the same name
are ever exactly the same.
The crystal structures of the main binary iron oxides and hydroxides are
illustrated in Fig. 11. We first present the structural and magnetic properties of
each in turn and then discuss structurally related binary oxides with other magnetic
cations. More complex, ternary oxides are treated in a following section.

Table 5 Magnetic properties of iron oxides and hydroxides


Space Lattice TC , TN σ
Mineral Ideal formula group parameters (K) Order (Am2 kg−1 )
Wüstite FeO Fm3m 431 198 AF <1
Magnetite Fe3 O4 Fd3m 840 853 FI 92
Hematite αFe2 O3 R3c 504; 1377 980 wF 0.5
Maghemite γFe2 O3 P41 835; 2499 950 FI 84
Goethite αFeOOH Pbnm 996; 302; 461 360-405 AF <1
Akaganénite βFeO(OH,cl) I2/m 1056; 303; 1048 295 AF –
90.0◦
Lepidocrocite γFeOOH Cm21 308; 1250; 387 70 AF <1
Feroxhyte δFeOOH P3ml 293; 460 455 Planar ≈10
AF
Ferrihydrite 5Fe2 O3. 9H2 O P63 mc 600; 910 poorly ∼100 AF ≈5
crystallized
Ferric gel Fe(OH)3 nH2 O – Amorphous 10 Sp ≈9
Amakinite (FeMg)OH2 p3m1 692; 1452 20 AF
AF, antiferromagnetic; Fi, ferrimagnetic; wF, weak ferromagnetic; Sp, speromagnetic
864 J. M. D. Coey

Fig. 11 Crystal structures of iron oxides (a) hematite, (b) magnetite, (c) maghemite, (d) wüstite
and hydroxides, (e) goethite, (f) akagaénite, (g) lepidocrocite, (h) feroxhyte, and (i) amakinite.
Key: Red O2- ; light brown Fe3+(oct) ; dark brown Fe3+(tet) ; green mixed iron valence; blue Fe2+ ;
violet H+ ; Green Cl- . In maghemite 1/6 of the octahedral sites are unoccupied

Hematite

The most abundant magnetic oxide is hematite (αFe2 O3 ). It is an insulating, rock-


forming mineral that is mined in vast quantities, especially in Australia and Brazil,
as a primary iron ore in formations known as itibarites. These ores, with ≥60 wt%
iron, can be used directly in the blast furnace. Lower-grade ores with ≈ 30 wt% iron
that require concentration are produced in large quantities in China. Normalized
annual iron ore production runs at about 2Gt/yr., with prices around 60$/t. Massive
hematite has a shiny black appearance, with a characteristic red streak. The powder
itself is dark red.
The crystal structure of hematite is that of corundum or sapphire (αAl2 O3 ).
It is based on a hexagonal close-packed oxygen array, where Fe3+ ions occupy
two-thirds of the octahedral sites; the other one-third are vacant. The structure is
rhombohedral, space group R3c with two formula units (Z = 2) in the unit cell. As
with other rhombohedral crystals, it is convenient to index the structure on a larger
hexagonal cell with Z = 6 and a = 504, c = 1375 (Fig. 11a), where iron occupies
the octahedral 12c site. The rhombohedral
111 direction becomes the hexagonal c-
axis. The density is 5260 kgm−3 . All iron sites are equivalent, and the structure may
17 Magnetic Oxides and Other Compounds 865

be regarded as an assemblage of edge- and corner-sharing FeO6 octahedra, with a


principal Fe3+ -O-Fe3+ superexchange bond angle of 132◦ .
Hematite is strongly antiferromagnetic, with a Néel temperature of 960 K, almost
the highest for any magnetic oxide [74]. The four ferric ions in the rhombohedral
unit cell are ordered + - -  + along the
111 direction, where + and – represent
the different magnetic sublattice orientations and  is a vacant site. Below a spin
reorientation transition at TM = 265 K, known as the Morin transition, the easy axis
lies at about 10 degrees ro the hexagonal c-axis [56]. At TM , the antiferromagnetic
axis changes, driven by a change in sign of the net magnetic anisotropy causing the
moments to rotate continuously by almost π/2 in a temperature interval of about
10 K into the basal plane [2, 74]. The in-plane moments are no longer strictly
collinear at room temperature, for example, because of a weak Dzyaloshinskii-
Moriya interaction, Eq. 6, where the interaction vector (≈ 0.1 K) must lie along
the c-axis for reasons of symmetry. The unit cell itself has no center of symmetry.
The average value of is ≈ 20 K and the slight canting of the antiferromagnetic
sublattices at an angle of order produce a weak resultant magnetization of 2.5 kA
m−1 or 0.014 μB /Fe. Although Fe3+ is an S-state ion with no orbital moment
and no magnetocrystalline anisotropy to first order, the 12c cation site has trigonal
symmetry (3), and the crystal field mixes some of the orbital character of the excited
states into the ground site, leading to weak single-ion magnetocrystalline anisotropy.
The magnetic dipole field at the Fe sites is of order 1 T (only in cubic lattices
does it sum to zero), giving a dipole-dipole anisotropy term. The two contributions
are delicately balanced with opposite sign, but their temperature dependences are
slightly different, varying as
Sz 2 and
Sz 2 , which explains the spin reorientation at
the Morin transition, where the sum changes sign [2], Fig. 12. The reversible Morin
transition established that the weak ferromagnetism of hematite was an intrinsic

Fig. 12 The Morin transition


in hematite. The cancellation
of the two contributions to the
anisotropy field occurs at the
Morin transition, where the
Néel vector (double blue
arrow) rotates from parallel to
the c-axis below TM = 260 K
to perpendicular to the c-axis
above TM . A weak
ferromagnetic moment then
appears due to spin canting in
the basal plane, produced by
the Dzyaloshinskii-Moriya
interaction. Ms is the weak
net moment TM
866 J. M. D. Coey

property associated with spin canting, and not parasitic ferromagnetism due to traces
of magnetite impurity. The anisotropy energy K1 at low temperatures is 9 kJm−3 ,
and hematite exhibits a spin flop transition in 6.9 T at 4 K. Because of the weakness
of the net magnetization, the anisotropy field μ0 Ha = 2 K1 /Ms is rather large and
powder samples often exhibit coercivity at room temperature.

Magnetite

Magnetite, Fe3 O4 , is the most famous magnetic mineral of them all and the other
major rock-forming iron oxide. It is a black ferrimagnetic conductor that is very
common in igneous rocks, and it was the main component of lodestones, the
original, natural permanent magnets that became magnetized by lightning strikes
on rocky outcrops. Ti-substituted magnetite is a common natural variant, and it
is our main source of information on paleomagnetism, the history of the Earth’s
magnetic field. Magnetite is important as iron ore, as a black pigment, toner (ink),
and in ferrofluids. Biogenic magnetite is produced by bacteria, pigeons, and other
creatures. The oxide can be a nuisance in sensitive magnetization measurements as
it is a widespread magnetic contaminant present in particulate air pollution.
The inverse spinel structure of magnetite (spinel is the mineral MgAl2 O4 )
is based on an face-centered cubic array with 32 oxygen ions, space group
Fd3m, Z = 8, where one-eighth of the tetrahedral interstices (8a, A-sites) are
occupied by ferric iron and one-half of the octahedral interstices (16d, B-sites)
are occupied by an equiatomic mixture of ferrous and ferric iron (Fig. 11b).
This cation distribution, with trivalent cations on A-sites, is known as the inverse
distribution. The cubic lattice parameter is a0 = 840 pm. Spinel itself has the
less common normal distribution, with the divalent Mg2+ cations on A-sites and
the trivalent Al3+ cations on B-sites. The chemical formula of magnetite can be
written as [Fe3+ ]{Fe2+ Fe3+ }O4 ; here [] denotes the A-sites, and {} denotes the
B-sites where there is rapid Fe2+ -Fe3+ electron exchange. The octahedral and
tetrahedral sublattices are coupled antiparallel by superexchange via 127◦ A-O-B
bonds, giving a ferrimagnetic structure with a net moment of about 4 μB and a room-
temperature magnetization of 92 Am2 kg−1 or 480 kAm−1 . The Curie temperature
is 860 K. The mobile electron associated with Fe2+ on the octahedral sites occupies
a one-sixth-filled minority-spin t2g band and the average B-site configuration is
(t2g 3 eg 2 )↑ (t2g 0.5 )↓ with
n = 5.5, so magnetite was expected to be half-metallic
with a spin gap in the majority density of states. There is limited evidence for an
exceptionally high-spin polarization at room temperature [81]. The 3d conduction
electrons interact electrostatically with the oxygen anions to form small polarons
with large effective mass, which move among the B-sites by thermally activated
hopping [117].
Magnetite undergoes the charge-ordering Verwey transition at 120 K [110]. This
is transition where the conductivity decreases abruptly by a factor ∼100 on cooling
and the symmetry of the crystal is lowered from cubic to monoclinic. Verwey’s
original model of regular charge ordering of Fe3+ and Fe2+ ions on octahedral sites
[108] was oversimplified, and a highly complex ordering of fractional charges in
17 Magnetic Oxides and Other Compounds 867

these sites described by the superposition of 168 atomic displacement waves with
iron trimer motifs has been established [92].
The magnetocrystalline anisotropy of magnetite K1c is −14 kJm−3 at room
temperature and K2c is −3 kJm−3 . However, K1c changes sign on cooling to
135 K, where the easy axis switches from
111 to
100 . The room-temperature
magnetostriction constants are λ111 = 78 10−6 and λ100 = −20 10−6 and the
polycrystalline average λs = 3/5λ111 + 2/5λ100 = 39 10−6 .

Maghemite

Maghemite, γFe2 O3 , is a brown ferric sesquioxide polymorph which forms


as a low-temperature oxidation or weathering product of magnetite, and it is
a common constituent of magnetic tropical soils (The prefix “sesqui” means
one and a half.). Maghenite has a spinel-type structure closely related to
that of magnetite where one-sixth of the B-sites are vacant, leading to a
formula [Fe3+ ]{Fe3+ 5/3 1/3 }O4 . The vacancies may be disordered (Fig. 11c)
or else adopt an ordered superstructure in a tetragonal unit cell with c = 3a
[51]. The oxide is ferrimagnetic, with a room-temperature magnetization
of 78 Am2 kg−1 or 380 kAm−1 and a density of 4860 kgm−3 . At low
temperatures, the magnetization rises to 82 Am2 kg−1 , a little less than the
88 Am2 kg−1 expected for a moment of 5μB per Fe3+ ion (2.5 μB per Fe2 O3
formula unit) because of the partly covalent bonding.
Maghemite converts to hematite, the more stable polymorph, when heated in
air at a temperature in the range 250–750 ◦ C that depends sensitively on sample
quality. The transition involves changing the packing of the layers in the close-
packed oxygen lattice from cubic to hexagonal. Because of the transition, the Curie
temperature of maghemite has to be extrapolated from measurements at lower
temperatures.
The average magnetostriction is λs = −9 10−6 and the cubic anisotropy constant
K1c = −5 kJm−3 .

Wüstite

The most reduced iron oxide is wüstite – FeO – a rare mineral that crystallizes
in the cubic halite (NaCl) structure, space group Fm3m with Z = 4 (Fig. 11d).
The lattice parameter is a0 ≈ 431 pm; ferrous iron occupies octahedral sites, but
the lattice is subject to a slight rhombohedral distortion to lower the energy of the
Fe2+ 3d6 cations. The oxide is generally nonstoichiometric, with an Fe/O ratio in the
range 0.88–0.96. Some ferric iron, needed for charge neutrality, occupies octahedral
sites and some tetrahedral sites in the halite structure that have nearby octahedral
vacancies [53].
The ferrous cation sites form a face-centered cubic lattice where antiferromag-
netic J1 interactions of a cation with its 12 nearest-neighbors are partly frustrated,
868 J. M. D. Coey

but the J2 interactions with the six second-neighbors are straightforwardly anti-
ferromagnetic, with 180◦ Fe-O-Fe bonds. The type II antiferromagnetic structure
is composed of alternating ferromagnetic (111) planes, with a Néel temperature
of 198 K. The direction of sublattice magnetization is
111 . The two exchange
parameters can be deduced from measurements of the Néel temperature TN and the
paramagnetic Curie temperature θ .

Iron Hydroxides

There is a series of iron hydroxides. The least hydrated are the four oxyhydroxides
with the chemical formula FeOOH, equivalent to Fe2 O3. H2 O. The former empha-
sizes the presence of OH− anions in the crystal structure.

Goethite and Other Oxyhydroxides


Goethite, αFeOOH, is the most common iron oxyhydroxide. It is a yellow-brown
orthorhombic ore mineral, an important constituent of soils and sediments, and
a major component of rust. It was named for Goethe, the German poet who
was an avid collector of minerals and Commissioner for Mines in the Duchy
of Saxe-Weimar. The crystal structure is presented in Fig. 11e. The material is
antiferromagnetic with a Néel temperature variously reported in the range 360–
405 K and a weak moment somehow due to uncompensated spins along the
orthorhombic b axis. However, the magnetocrystalline anisotropy is very low, ≈
0.2 kJm−3 at room temperature, and this small value, combined with a tendency for
the samples to form fine crystallites, often leads to superparamagnetic behavior. A
spin flop observed in 20 T at 5 K when the field is applied along the b axis [26]
would appear at lower fields, were it not for the weak moment along b. Goethite is
dehydrated by heating to 250–400 ◦ C, to form hematite.
Akaganénite βFeO(OH,Cl) is an iron oxyhydroxide that incorporates some Cl−
ions to stabilize the monoclinic structure (Fig. 11f). It is antiferromagnetic below
295 K.
Lepidocrocite, γFeOOH, is an antiferromagnet with a low Néel temperature of
70 K; it is a minor component of some soils and sediments.
δFeOOH, feroxyhyte is the fourth crystalline ferric oxyhydroxide. It has the
greatest density, and unlike others, it is a planar antiferromagnet with an hexagonal
structure of ferromagnetic c-planes (Fig. 11h). Better-crystallized specimens order
magnetically below 455 K, with a small net magnetization of about 10 Am2 kg−1
due to imperfect compensation of the antiferromagnetic sublattices [6].

Ferrihydrite and Ferric Gel


Poorly crystallized hydrated ferric material is very common at the Earth’s surface.
Often formed by oxidation of water rich in dissolved ferrous iron, it appears in
soils and sediments and in colloidal form in surface water. The material also serves
as a precursor for the more stable ferric oxyhydroxides mentioned above that are
also produced at the surface by weathering or during pedogenesis. The material
17 Magnetic Oxides and Other Compounds 869

exhibits a varying degree of crystallinity and hydration [91]. Three categories are
distinguished, a “six-line” ferrihydrite, a “two-line” ferrihydrite, and an essentially
amorphous ferric gel. Ferrihydrite is a poorly crystallized hydrated ferric oxide
with notional composition 5Fe2 O3. 9H2 O. The better-crystallized variety shows
six broad X-ray reflections and orders magnetically below 100 K with a small
uncompensated moment. The two-line version shows just two very broad reflections
corresponding to lattice spacings of 250 and 150 pm and behaves similarly. The
ferric gel exhibits no clear Bragg reflections. Below 10 K, spins freeze in a
speromagnetic structure with antiferromagnetic nearest-neighbor correlations but
essentially random orientations of more distant neighbors leading to a small net
magnetization [24]. Superparamagnetic behavior is observed up to about 100 K.
The structures can be modeled as a network of edge- or corner-sharing octahedra.
A related synthetic amorphous magnet is FeF3 , where the structure is an
octahedral random network of corner-sharing FeF6 octahedra and the cations form
odd- and even-membered rings. The spins freeze below 30 K in a speromagnetic
spin structure, although the pairwise antiferromagnetic superexchange bonds (frus-
trated in the odd rings) would, if unfrustrated, give much higher antiferromagnetic
ordering temperature.

Ferrous Hydroxide
The natural ferrous hydroxide amakinite Fe(OH)2 has the hexagonal CdI2 structure
of brucite Mg(OH)2 and pyrochorite Mn(OH)2 . It is a layered structure based on a
hexagonal close-packed arrangement of OH ions, where the divalent cations fill all
the octahedral sites in alternate layers. Amakinite orders magnetically at 34 K, in
a planar antiferromagnetic structure of ferromagnetic planes. Green rust is a poorly
crystallized variant that is unstable in air. The Néel temperature of pyrochorite is
12 K, but the planes in that case have a spiral magnetic structure.

Ferrites

There are some very important families of synthetic stoichiometric insulating oxides
of ferric iron and another cation, known collectively as ferrites. These include
compounds with the spinel, garnet, orthoferrite, and hexaferrite structures. A few
natural examples exist, but ferrites are generally pure synthetic materials, prepared
by standard ceramic techniques. The vast majority of bulk magnetic oxides used for
practical purposes are hard or soft ferrites. The subject was first summarized by Smit
and Wijn [96] and other books include Valenzuela [104] and Dionne [38]. Many of
the original numerical data are to be found in Landholdt-Börnstein [9].

Spinels

The family with the cubic spinel structure, space group Fd3m, includes ZnFe2 O4 ,
MgFe2 O4 , MnFe2 O4 , Fe3 O4 , CoFe2 O, NiFe2 O4 , and Li0.5 Fe2.5 O4 . Only the first of
870 J. M. D. Coey

these has the normal cation distribution with iron on octahedral B-sites and zinc on
tetrahedral A-sites, whereas the others are inverse with Fe3+ on tetrahedral sites,
which leads to high-temperature ferrimagnetic ordering. Lithium ferrite has the
highest Curie temperature (also known as the ferrimagnetic Néel temperature) of
943 K. The extrapolated Curie temperature of maghemite is even higher, but it is
unattainable because γFe2 O3 converts to αFe2 O3 first. Cobalt ferrite has by far the
greatest magnetocrystalline anisotropy and magnetostriction, with [100] easy axes,
whereas all the others have [111] easy axes. These outstanding values arise from an
inverse cation distribution with Co2+ 3d7 ions on B-sites, which have trigonal 3m
point symmetry. These B-site ions are in a high-spin state t2g 5 eg 2 with S = 3/2 and a
spin moment of 3 μB . They can also have a significant unquenched orbital moment
of ∼0.6 μB , which is responsible for the strong cubic anisotropy K1c ≈ 290 kJm−3 .
Although the moment of an isolated Co2+ ion aligns along a local
111 trigonal
axis, the resultant bulk anisotropy lies along
100 [63]. The net moment is about
3.6 μB per formula at room temperature, and the magnetization of bulk samples is
455 kAm−1 or 86 Am2 kg−1 (455 emu/g or 86 emu/g), based on the X-ray density
of 5290 kgm−3 . The ferrimagnetic Néel temperature of CoFe2 O4 is 790 K, so the
magnetization values at T ≈ 0 are slightly higher.
Another consequence of the unquenched orbital moment on the Co2+ is an
exceptionally large magnetostriction. Values of λ100 for CoFe2 O4 crystals range
from −250 to −650 10−6 , depending on their precise composition. The value
of the other cubic magnetostriction constant λ111 is about +170 10−6 , and the
polycrystalline averages λs = (2λ100 + 3λ111 )/5 are negative, but quite variable.
A consequence of the magnetostriction is a slight distortion of the unit cell from
cubic Fd3m to tetragonal I41 /amd, with a tetragonality (a – c)/c ≈ 0.1%.
Copper ferrite, CuFe2 O4 , has a mostly inverse cation distribution that depends
on the preparation procedure. At temperatures below 700 ◦ C, the spinel structure is
tetragonally distorted with c/a = 1.06, but it is cubic when prepared above 800 ◦ C.
The anisotropy changes sign at 350 K, and there is a complex sequence of magnetic
phase transitions below room temperature.
Zinc ferrite has the normal cation distribution. The only magnetic species is
Fe3+ on octahedral sites, which form a network of corner-sharing tetrahedra in
the spinel structure. When the B-B interactions are antiferromagnetic, as they are
in ZnFe2 O4 , the exchange bonds are fully frustrated, which explains the low-spin
freezing temperature Tf , < 10 K. Magnetic properties of spinel ferrites are discussed
in detail in references [10, 64], and salient data on ferrimagnetic end-members are
summarized in Table 6. Mineral names are included, where they exist.
A related series of ferrites with formula FeM2 O4 are of less interest magnetically.
Chromite, FeCr2 O4 , is the principal natural source of chromium for stainless steel.
It orders magnetically at about 70 K, with a small magnetization of ∼0.08 MAm−1 .
Ulvöspinel, TiFe2 O4 , is an inverse spinel with a0 = 840 nm with formal
valences Ti4+ and Fe2+ . It orders antiferromagnetically at 120 K. There is a huge
range of natural and synthetic solid solutions with the spinel structure. Natural
titanomagnetites Fe3-x Tix O4, solid solutions between magnetite and ulvöspinel with
x ∼ 0.6, are important for palaeomagnetism, because fine grains of this mineral
17 Magnetic Oxides and Other Compounds 871

Table 6 Magnetic properties of spinel ferrites


a0 Ms K1 λs
Oxide Mineral (pm) TC , (K) (MAm−1 ) (kJm−3 ) (10−6 ) ρ( m)
MgFe2 O4 Magnesioferrite I 830 713 0.18 −3 −6 105
Li0.5 Fe2.5 O4 – 829 943 0.33 −8 −8 1
CrFe2 O4 I 834 430 0.18 60
MnFe2 O4 Jacobsite I 852 575 0.50 −3 −5 105
Fe3 O4 Magnetite I 840 853 0.48 −13 40 10−1
CoFe2 O4 I 839; 790 0.45 290 −110 105
838
NiFe2 O4 Trevorite I 834 865 0.33 −7 −25 102
CuFe2 O4 I 843 710 0.19 −3 −3 10
ZnFe2 O4 N 844 Tf ∼ 9 K – 1
γFe2 O3 Maghemite – 834 985 0.43 −5 −5 ∼1
I, inverse; N, normal

are the main carriers of natural remanence in rapidly cooled basaltic lavas used for
paleomagnetic studies [39]. Titanohematites, solid solutions between hematite and
ilmenite FeTiO3 , are also of interest in this context.
Optimized nickel-zinc ferrite and manganese-zinc ferrite are widely used as
cores in high-frequency inductors. The properties can be tuned by tailoring the
composition, microstructure, and porosity. Manganese-zinc ferrites offer initial
permeabilities of up to 10,000 and low coercivity; they may be used at frequencies
up to about 1 MHz. Nickel-zinc ferrites have lower permeability, generally <1000,
and lower magnetization but high resistivity ρ of up to 106 m, which allows them
to be used at radiofrequencies of 300 MHz or more.

Garnets

Garnets are oxides with formula A3 B2 C3 O12 that crystallize in the Ia3d space
group with a large cubic cell, Z = 8, having three cation sites 24c, 16a, and 24d
with eightfold, sixfold, and fourfold oxygen coordination. Natural garnets include
pyrope Mg3 Al2 Si3 O12 , grossular Ca3 Al2 Si3 O12 , andradite Ca3 Fe3+ 2 Si3 O12 , and
almandine Fe2+ 3 Al2 Si3 O12 . The latter two order magnetically below 10 K. Some
garnets are red or green gemstones.
The garnets are well known in magnetism because of the series of R3 Fe5 O12
compounds that adopt this structure when R3+ is yttrium or a tripositive 4f ion
from Sm to Yb. Yttrium iron garnet Y3 Fe5 O12 (commonly known as YIG) is a
green ferrimagnetic insulator with a bandgap of 2.8 eV and a Curie temperature of
560 K. The net uncompensated 24d-site moment leads to a room-temperature mag-
netization of 27.6 Am2 kg−1 or Ms = 144 kAm−1 . Anisotropy (K1 = −2 kJm−3 )
and magnetostriction (λ100 = 1 ppm, λ111 = −3 ppm) are both very small,
which ensures outstanding radiofrequency magnetic properties and a very narrow
872 J. M. D. Coey

Table 7 Magnetic properties of rare-earth iron garnets


Y Sm Eu Gd Tb Dy Ho Er Tm Yb Lu
a0 (pm) 1238 1253 1250 1247 1244 1241 1238 1235 1232 1230 1228
M (kAm−1 ) 141 141 94 10 25 35 66 92 114 124 145
M(0) (μB ) 5.0 5.4 2.8 16.0 18.2 16.9 15.2 10.2 1.2 0.0 5.1
TC (K) 553 578 566 564 568 563 567 556 549 548 549
Tcomp (K) – – – 286 246 226 137 83 – ∼0 –

ferromagnetic resonance linewidth H ∼ 40 Am−1 . YIG is widely used for


microwave components such as circulators and filters. It also has good magneto-
optic properties when doped with Bi. Thin films find applications in spintronics.
Magnetic properties of YIG and the series of rare-earth garnets are discussed
in detail in reference [47], and salient data are summarized on Table 7. Weak
superexchange coupling of the rare-earth and iron sublattices leads to parallel or
antiparallel alignment of the net moments for light and heavy rare earths, respec-
tively. For example, Gd3 Fe5 O12 has a low-temperature moment of approximately
3 × 7 − (3 × 5 − 2 × 5) = 16μB per formula unit. However, the Gd sublattice
moment falls off much more rapidly with temperature than that of the Fe sublattices,
which leads to a sign change of the net moment at a compensation temperature Tcomp
of 286 K. The temperature dependence of the net moments is illustrated in Fig. 13.
Compensation temperatures range from 286 K for Gd to ∼0 K for Yb. Despite
the small anisotropy, the anisotropy field Ha = 2 K1 /Ms diverges at compensation,
which leads to a large coercivity in the vicinity of Tcomp .

Orthoferrites

The orthoferrites are rare-earth iron sesquioxides RFeO3 with an orthorhombically


distorted (quasi-tetragonal) perovskite structure with space group Pbnm. The unit
cell contains four formula units. Other perovskites are presented in Sect. “Ferrites”
Data on the RFeO3 compounds, which form with yttrium and all the 4f series, are
listed in Table 8 [103]. The oxides are antiferromagnetic, with Néel temperatures in
the range 623–740 K that scale with the cosine of the Fe-O-Fe superexchange bond
angle. In most of the compounds, the antiferromagnetic axis is perpendicular to the
c-axis, and the Dzyaloshinski-Moriya interaction causes the moments to become
slightly canted in the c-plane, like in henatite.
Orthoferrite films grown by liquid-phase epitaxy were the basis of a magnetic
bubble memory with no moving parts in the 1970s [42]. The oxides are transparent,
with a large Faraday rotation, which makes them interesting for optical sensing
and actuation. It is possible to excite and observe very-high-frequency terahertz
antiferromagnetic resonance modes with short laser pulses [72].
17 Magnetic Oxides and Other Compounds 873

Fig. 13 Temperature dependence of the magnetic moment per formula unit of some rare-earth
garnets

Table 8 Magnetic properties of rare-earth orthoferrites


Y La Pr Nd Sm Eu Gd Tb Dy Ho Er Tm Yb Lu
a (pm) 528 556 549 544 539 537 535 533 530 528 526 525 523 521
b (pm) 559 557 558 557 559 561 562 560 560 559 558 558 556 555
c (pm) 760 786 781 775 771 769 767 764 762 760 759 758 757 757
M (kAm−1 ) 7.4 7.2 – – 6.5 6.4 7.9 6.3 6.6 6.8 6.8 6.7 7.6 9.2
M (mμB /fu) 45 47 – – 41 40 49 39 40 41 41 40 45 54
TN (K) 643 740 709 690 674 664 663 650 646 643 640 631 631 623

Hexagonal Ferrites

Hexagonal ferrites, an important family of magnetic iron oxides, structurally related


to the minerals magnetoplumbite (PbFe12 O19 ) and hibonite (CaAl12 O19 ), were
developed at Philips in the Netherlands around 1950 and discussed by Smit and Wijn
[96]. These were the first mass-market permanent magnets that could be fabricated
in any desired shape without self-demagnetization. The compounds, known as
874 J. M. D. Coey

Fig. 14 (a) The magnetic structure of SrF12 O19 , showing the 2b bipyramidal site, (b) the principal
superexchange paths, and (c) the principal contributions to the uniaxial anisotropy

M – ferrites, have composition AFe12 O19 , where A is a large divalent cation [63].
Processing routes for both sintered and polymer-bonded hexagonal ferrites are well-
established [97]. Both BaFe12 O19 and SrFe12 O19 are produced in huge quantities,
of order a million tons per annum corresponding to more than 100 g on average for
everyone on Earth.
The magnetoplumbite structure is illustrated in Fig. 14a). It can be regarded as an
hexagonal packing of O2− and A2+ ions in a tall unit cell with space group p63 /mmc
having a ≈ 590 pm, c ≈ 2300 pm, and Z = 2. There are five types of interstitial sites
for ferric ions, three with octahedral symmetry (12k, 4f2 , and 2a), one tetrahedral
(4f1 ), and one fivefold coordinated trigonal bipyramidal site (2b). The structure
can be regarded as a stacking of alternate (111) slices of spinel (S blocks, Fe6 O8
containing 2a, and half the 12k octahedral sites and the 4f1 tetrahedral sites) and
triple layers of hexagonal close-packed oxygen (R blocks, AFe6 O11 , containing the
2b sites, and the remaining octahedral sites), where one in four of the oxygen in the
middle layer is replaced by A2+ . The stacking in the unit cell is SRS*R*, where *
indicates rotation by π around the hexagonal c-axis. A large family of related hexag-
onal oxides is obtained by different stackings of R, S, and T = A2 Fe8 O14 blocks.
The main superexchange paths are indicated on Fig. 14b), together with the
values of the exchange constants [94]. The strongest couplings tend to coincide with
the straightest bond angles. The ferrimagnetic structure is 12 k↑;2a↑;2b↑;4f1 ↓;4f2 ↓.
Since Fe3+ has a spin moment of 5μB and there are Z = 2 formula units in the unit
17 Magnetic Oxides and Other Compounds 875

Table 9 Properties of hexagonal ferrites at room temperature. (m0 is the moment at T = 0)


mo Ms K1
Ferrite a (pm) c (pm) TC (K) (μB /fu) (kAm−1 ) (MJm−1 )
BaM BaFe12 O19 589 2320 740 19.9 380 0.33
SrM SrFe12 O19 589 2304 746 20.2 370 0.35
PbM PbFe12 O19 590 2309 725 19.6 330 0.25
BaW BaM2 Fe16 O27 588 3250 728 27.6 410 0.30
BaX Ba2 M2 Fe24 O46 588 5570 735 47.5 280 0.30
BaY Ba2 M2 Fe12 O22 (M = Co.) 590 4360 613 9.8 185 −0.26
BaZ Ba2 M2 Fe24 O41 (M = Co.) 588 5230 683 31.2 267 −0.18

cell, the net moment per formula unit at T ≈ 0 is (8 – 4) × 5μB = 20μB . The net
exchange coupling of 12k iron to its neighbors is weakest, and this iron moment has
the strongest temperature dependence (Fig. 14c).
The M, W, and X ferrites have easy c-axis anisotropy. The Y and Z ferrites are
easy-plane when M = Co. The magnetic properties are presented in detail in [63]
and summarized in Table 9. Properties of the three M-ferrites are quite similar.
The anisotropy field at room temperature is around 1.8 T, more than three times
the polarization μ0 Ms . Cobalt-substituted Y and Z ferrites can exhibit very little
coercivity, and they are useful for microwave applications.
Strong magnetic anisotropy is not normally expected from Fe3+ 3d5 ions, which
have a 6 S ground state with a spherical half-filled shell. However, in highly distorted
sites, the crystal field can mix in nonspherical 4 F excited states, leading to an
effective spin Hamiltonian DSz 2 ; the 2b trigonal bipyramidal sites and 12 k distorted
octahedral sites in M ferrites contribute about half of K1 , and there is a contribution
of opposite sign from the anisotropic magnetic dipole field (Fig. 17c) [94]. The iron
on 2b sites is hopping rapidly between two the tetrahedral positions that make up
the bipyramid [78].
A long search for cation substitutions to increase the magnetization or anisotropy
of M ferrites has led to a slight improvement by replacing some Sr2+ + Fe3+ by
La3+ + Co2+ . The magnetization Ms in oxide ferrimagnets is unavoidably limited
by their partly compensated ferrimagnetic structure and the fact that most of the
volume is composed of nonmagnetic oxygen ions. These drawbacks are partially
compensated by the fact that larger moments are available from ionic cations than
from metallic atoms. An extensive review [89] discusses cation substitutions in
hexagonal ferrites and the synthesis and ceramic processing of these materials.
Crystallite orientation is a critical step for easy axis (hard magnet) or easy-plane
(microwave ferrite) materials.

Other Magnetic Oxides

This section discusses magnetic oxides where the magnetic cation is not iron. The
materials are generally synthetic, not natural in origin.
876 J. M. D. Coey

3d Oxides

The 3d cations tend to be trivalent or quadrivalent at the beginning of the series


and divalent or trivalent at the end. Hence, we find dioxides, sesquioxides, and
monoxides.

Monoxides
There is a series of isostructural antiferromagnetic insulating monoxides MnO, FeO,
CoO, and NiO with the NaCl structure (Fg 1d), which all have magnetic structures
of alternating ferromagnetic (111) planes. MgO is a nonmagnetic analogue. MnO
(manganosite) was the first material where antiferromagnetic order was identified,
from a series of magnetic Bragg neutron reflections that appeared below the Néel
temperature of 118 K. CoO, with high-spin Co2+ on the octahedral sites, has its
Néel temperature close to room temperature. The lattice has a slight tetragonal
distortion. NiO has the highest Néel temperature of 525 K, and the structure, like
that of FeO (wüstite), has a slight trigonal distortion. NiO is a green charge-transfer
insulator with a bandgap of 4 eV when stoichiometric [65], but it becomes a black
p-type polaron conductor when cation-deficient due to holes in the oxygen 2p band.
Stoichiometric NiO was used for exchange bias of early spin valves.
The magnetic properties of the cubic antiferromagnetic monoxides, including
wüstite, are summarized in Table 10. The nearest-neighbor superexchange inter-
action J1 with the 12 nearest-neighbor cations on an fcc lattice is frustrated, but
the interaction J2 with the six second-neighbor cations which lie on a simple cubic
lattice, with a 180◦ M-O-M superexchange bond, favors antiferromagnetic order.
CuO (tenorite) has a monoclinic structure with space group C2/c and a = 469 pm,
b = 343 pm, c = 5i3 pm, and β = 99.5◦ It orders magnetically at 230 K in a helical
structure, but it becomes a commensurate antiferromagnet below 213 K, where it is
also ferroelectric. It is the first binary multiferroic [111].

Sesquioxides
Ti, V, and Cr form sesquioxides which, like hematite, crystallize in the corundum
structure. Ti2 O3 is diamagnetic, with a small bandgap of only 0.2 eV between
the lower and upper Hubbard bands. The Ti3+ 3d1 electrons are thought to form
pair bonds, rather than ordering antiferromagnetically. V2 O3 exhibits a spectac-
ular first-order insulator-metal transition at the Néel temperature (150 K), where

Table 10 Magnetic properties of cubic antiferromagnetic monoxides


Oxide Ion Spin Lattice parameter (pm) TN (K) θ (K) J1 (K) J2 (K)
MnO 3d5 5/2 445 118 −610 −7.2 −3.5
FeO 3d6 2 431 R 198 −570 −7.8 −8.2
CoO 3d7 3/2 426 T 291 −330 −6.9 −21.2
NiO 3d8 1 418 R 525 −1310 −8.0 −110
R, rhombohedral distortion; T, tetragonal distortion
17 Magnetic Oxides and Other Compounds 877

Table 11 Magnetic properties of sesquioxides with the corundum, spinel, or bixbyite structures
Mineral Lattice
Oxide name parameters (pm) TN (K) Magnetism Other information
Ti2 O3 Tistarite R3c 514; 1366 – Diamagnetic p-type,
gap = 0.2 eV
V2 O3 Karelianite R3c 503; 1362 150 Antiferro Monoclinic below
TN
Metal-insulator
transition
Cr2 O3 Eskolaite R3c 496; 1359 306 Antiferro Magnetoelectric
αMn2 O3 Bixbyite Ia3 941 80 Antiferro First-order
transition at 25 K;
stabilized by iron
γMn2 O3 579; 940 39 Ferrimagnet
αFe2 O3 Hematite R3c 504; 1375 960 Canted af Morin transition at
265 K
γFe2 O3 Maghemite P41 32 835 985 Ferrimagnet Converts to
hematite on heating

the conductivity can change by as much as ten orders of magnitude [75]. The
low-temperature structure is monoclinic. Cr2 O3 has a different antiferromagnetic
stacking + − + − below its Néel point TN = 306 K compared to +– – +
for αFe2 O3 , and there is no spin reorientation. The signs refer to the direction of
magnetization of the four magnetic sublattices marked by the four cations along the
c-axis in Fig. 11a). Cr2 O3 is weakly magnetoelectric; a single antiferromagnetic
domain state produced by cooling below TN in parallel electric and magnetic
fields exhibits a small electrically induced magnetic moment. The effect is caused
by relative displacements of Cr3+ and O2− , in the electric field, which slightly
modifies the crystal field. The magnitude of the magnetoelectric coefficient α = μ0
M/E ≈ 1psm−1 . Mn2 O3 can crystallize in two different cubic structures; one is the
bixbyite structure, and the other is a tetragonal structure related to spinel. Properties
of these compounds are summarized in Table 11.
Related oxides have an ordered arrangement of different metal cations on the
cation sublattices. These include a series of titanates MTiO3 , which are low-
temperature antiferromagnets; an example is ilmenite, FeTiO3 , where the formal
cation valences are Fe2+ and Ti4+ . Only the former has a magnetic moment, and
the compound orders antiferromagnetically at 40 K.

Dioxides
A series of 3d dioxides crystallize in the rutile structure. TiO2 itself exhibits
temperature-independent paramagnetism, but it does not order magnetically. VO2 ,
a material of potential interest for oxide electronics, exhibits a metal-insulator
transition at 341 K. It has the rutile structure above the transition and a distorted,
monoclinic structure below, but it too is not thought to order magnetically. CrO2 is
the only example we have of a ferromagnetic binary metal oxide. It is half-metallic,
878 J. M. D. Coey

Table 12 Magnetic properties of dioxides with the rutile structure


Oxide Ion Spin Lattice parameters (pm) Magnetism Tc (K)
TiO2 3d0 0 458;295 Van Vleck paramagnet –
VO2 3d1 1/2 455; 285 No magnetic order –
CrO2 3d2 1 442; 292 Ferromagnetic 390
βMnO2 3d3 3/2 440; 287 Antiferromagnetic 94

Table 13 Magnetic properties of mixed-valence oxides


Lattice
Oxide Mineral name Space group parameters Tc (K) Magnetism Other information
Mn3 O4 Hausmannite I41 /amd 577; 944 41 Canted Tetragonally
ferrimagnet distorted spinel,
normal distribution
Fe3 O4 Magnetite Fd3m 940 853 Ferrimagnet Verwey transition
at 119 K
Inverse distribution
Co3 O4 – Fd3m 808 40 Antiferromagnet Low-spin cobalt,
normal distribution

with a gap in the minority-spin density of states. CrO2 can be grown in the form
of 300 nm acicular nanoparticles with an aspect ratio of 8:1. It was widely used
for magnetic videotapes, which could be easily duplicated by thermal imprinting,
thanks to the convenient Curie temperature at TC = 390 K. It cannot be increased
by cation substitution. MnO2 is an antiferromagnetic insulator with a low Néel
temperature (Table 12).

Mixed-Valence Oxides
The mixed-valence manganese oxide Mn3 O4 (Hausmannite) is a spinel-type oxide
containing a mixture of Mn2+ 3d5 and Mn3+ 3d4 ions. Here the crystal field
stabilization energy favors B-sites for Mn3+ and a normal cation distribution. There
is no electron hopping, and the unit cell with Z = 4 is tetragonally distorted. The
oxide orders ferrimagnetically at 41 K, with [010] as the easy axes, and there is
a noncollinear, canted magnetic structure driven by local anisotropy at the Jahn-
Teller-distorted B-sites.
Co3 O4 has an undistorted cubic spinel structure with a normal cation distribution
and a0 = 808 pm. The Co3+ 3d6 on octahedral sites is in a low-spin state, where
it has no moment, and the Co2+ 3d7 is also in a low-spin state. The oxide orders
antiferromagnetically at 40 K (Table 13).

Perovskites

A large family of magnetic oxides ABO3 have structures related to cubic perovskite,
CaTiO3 . Here A is often a rare-earth or alkaline earth cation coordinated by 12
17 Magnetic Oxides and Other Compounds 879

Fig. 15 The cubic perovskite ABO3 and double perovskite A2 BB’O6 structures

oxygen anions, and B is a transition-metal ion coordinated by an oxygen octahedron


(Fig. 15a).
A tolerance factor t is defined in terms of the ionic radii as:

t = (rA + rO ) / 2 (rB + rO ) (7)

The factor t is 1 in the ideal structure, where the lattice parameter


a0 = 2(rB + rO ). When t > 1, the structure may be tetragonal or hexagonal, but the
oxides are cubic when 0.9 ≤ t ≤ 1.0 and when 0.7 ≤ t ≤ 0.9 a distorted version of
the structure with orthorhombic or rhombohedral symmetry forms that has a larger
unit cell, with tilted, twisted, or distorted BO6 octahedra. The orthorhombic Pbnm
√ √
or Pnma cells have dimensions ∼ 2a0 × 2a0 × 2a0 , with Z = 4, where a0 is
the lattice parameter of the simple cubic cell, Fig. 14a). The rhombohedral R3c

cell has a = 2a0 and α ≈ 60◦ with Z = 2, but it is often indexed on a bigger
√ √
hexagonal cell with a = 2a0 and c = 2 3a0 and Z = 6. When A is a trivalent
ion, such as a rare earth, the transition-metal B is also trivalent, but if A is a divalent
ion such as Ca2+ or Sr2+ , the transition metal should be formally quadrivalent.
The principal exchange interaction is antiferromagnetic B-O-B superexchange,
with a near-180◦ bond angle. A mixture of divalent and trivalent A-site ions leads
to mixed trivalent and quadrivalent B ions, as in the mixed-valence manganites
(La1-x Srx )MnO3 [27]. Electron hopping with spin memory is then possible among
the transition-metal cations, leading to conductivity, double exchange, and possible
ferromagnetic order.

Rare-Earth Orthoferrites and Related Compounds


These are perovskites where A is a trivalent rare-earth ion R3+ and B is a trivalent
3d transition-metal ion, which form in some cases across the whole rare-earth series.
The iron compounds introduced in Sect. “Wüstite” are a good example (Table 8).
880 J. M. D. Coey

Usually, these compounds are orthorhombic with space group Pbnm. The R-O-R
superexchange is very weak, as can be seen from the Néel temperatures of the rare-
earth sesquioxides in Table 19, which amount to only a few kelvins.
The orthotitanates with Ti3+ 3d1 are all Mott insulators; t is small, and there
are large rotations of the TiO6 octahedra. Compounds with a heavy rare earth or
yttrium exhibit orbital ordering of the occupied dyz and dzx orbitals, and they are
ferromagnetic; TC = 30 K for YTiO3 , for example, with a moment of 0.83 μB per
formula. The compounds of the light rare earths, with larger tolerance factors, are
simple G-type antiferromagnets where each Ti is antiparallel to all six Ti neighbors,
with a moment of 0.45 μB and TN = 146 K for LaTiO3 , for example; orbital order
is less pronounced.
The rare-earth orthovanadates form across the whole series from La to Lu,
with Néel temperatures in the range 100–140 K, decreasing with decreasing lattice
parameter and superexchange bond angles. Orbital order of the dyz and dzx orbitals
of the V3+ t2g 2 ions sets in at a temperature that is 10–70 K greater. The heavy
rare-earth ions order magnetically at temperatures of about 10 K or less.
Néel temperatures for the orthochromites are higher, 110–280 K, but the trend is
similar. There is no orbital order for Cr3+ ions, which have a half-filled t2g shell.
The rare-earth manganites are different. They involve the strong Jahn-Teller Mn3+
t2g 3 eg 1 ion and can crystallize either in the orthorhombic perovskite structure, with
a noncommensurate antiferromagnetic structure and a Néel point at about 40 K and
rare-earth ordering below 10 K, or else in a high-temperature hexagonal structure
with space group P63 cm, for yttrium and heavy rare earths where the Mn and R
cations are coordinated by five or seven oxygen anions, respectively. The Mn3+
ions form trimers, and the exchange is strongly frustrated, but antiferromagnetism
sets in at 70–110 K, with a spin liquid at higher temperatures. The compounds are
magnetoelectric and multiferroic, with a ferroelectric transition at 800–900 K.
The rare-earth orthoferrites with space group Pbnm (Table 8) have the highest
Néel temperatures, which decrease toward the end of the series with the lanthanide
contraction, similarly to the V and Cr perovskite series. Here the variation is from
740 K for La to 623 K for Lu, reflecting the decrease of the Fe-O-Fe superexchange
bond angle. They all have canted antiferromagnetic structures with a weak net
magnetization of order 10 kAm−1 , due to the D-M interaction, Eq. 6. The canting
angle is about 0.5◦ . The orthoferrites used to be of interest as materials for bubble
memory [42]. They exhibit fast spin dynamics associated with the antiferromagnetic
resonance mode and very high domain-wall velocities. They are also of interest as
potential ferroelectrics.
The cobalt in the RCoO3 series is nonmagnetic on account of the low-spin Co3+
state 3d6 t2g 6 in octahedral coordination.
All the nickelates, with the exception of LaNiO3 , which is a rhombohedral R3c
metal, have an insulating antiferromagnetic ground state and undergo a metal-
insulation transition at higher temperature. For Pr and Nd, where the transition is
first-order, this coincides with the Néel temperature but the transitions are distinct
for the others, as shown in Fig. 16. Néel temperatures range from 130 to 225 K. In
the insulating state, the orthorhombic structure undergoes a monoclinic distortion
17 Magnetic Oxides and Other Compounds 881

Fig. 16 Phase diagram of the rare-earth nickelates RNiO3 , plotted as a function of tolerance
factor, showing also the average Ni-O-Ni bond angle (after [16])

to P21/ n symmetry. The nickelates are regarded as charge-transfer insulators, and


they become metallic as rotation of the octahedra increases the 3d – 2p overlap and
bandwidth. The Ni-O-Ni bond angle mainly determines the physical properties [16]
(Table 14).
A related perovskite is BiFeO3 , which has the rhombohedral structure with
hexagonal lattice parameters a = 558 pm, c = 1387 pm. The material is a
ferroelectric with a ferroelectric Curie temperature of 1100 K and a polarization
of 0.9 Cm−2 . It orders additionally as a G-type antiferromagnet at TN = 653 K,
and it is one of the rare multiferroics to exhibit magnetic and ferroelectric order
simultaneously above room temperature. The antiferromagnetism is slightly canted,
with a weak cycloidal moment with period 62 nm.

Perovskite Solid Solutions


As with other oxide families, there is enormous scope to modify the magnetic
and electronic properties of perovskites in solid solutions, substituting cations of
different size and valence for those in the end-member compound. Changing the
tolerance factor, structure, electron occupancy, charge, and orbital order generates
rich and complex phase diagrams with abrupt or tuneable changes of properties
as a function of temperature or composition. The mixed-valence manganites
(R1-x Ax )MnO3 are probably the most famous examples, where R is a trivalent rare-
882

Table 14 Orthorhombic unit cell volumes v (pm3 ) and magnetic ordering temperatures (K) of RMO3 compounds
M|R Y La Ce Pr Nd Sm Eu Gd Tb Dy Ho Er Tm Yb Lu
V v 224 241 239 237 235 232 229 227 226 223 223 221 220 218
TN 114 143 135 142 139 133 123 111 110 112 110 101 102
Cr v 214 233 232 230 228 223 221 220 217 215 214 213 211 210 209
TN 141 282 257 239 224 193 181 170 158 146 141 133 124 118 112
Mn v 225 243 240 239 233 232 233 229 228 226 224 223 222 221
TN 42 139 95 85 47 43 46 42 41
Hex TN 72 57 76 80 82 88 90
Fe v 225 243 239 234 233 232 230 228 226 224 223 222 220 219
TN 643 740 709 690 674 664 663 650 645 643 640 631 631 623
Ni v 212 222 221 219 218 217 214 212 210 206 207 207
TN 145 130 200 225 205 180 163 148 144 144 138 130
J. M. D. Coey
17 Magnetic Oxides and Other Compounds 883

earth cation and A is a divalent alkaline earth ion, Ca or Sr [27]. We comment briefly
on the magnetic aspects here.
There are four basic types of magnetic order in perovskites: the F, A, C, and G
types. F is ferromagnetic, A is a planar antiferromagnet composed of ferromagnetic
planes stacked antiferromagnetically, C is composed of ferromagnetic chains
antiferromagnetically aligned, and G is a straightforward antiferromagnet where
all the magnetic nearest-neighbors of a magnetic ion are aligned antiparallel to
it. Symmetry-allowed magnetic modes may combine different types of magnetic
order, with the moments along the principal axes. For example, Gz may combine
with Fx . The end-member LaMnO3 is an A-type antiferromagnetic insulator, with
TN = 150 K. On doping with Sr or Ca, the oxides first become í-type insulators
and then F-type metals. Optimal doping is at x ≈ 0.3, where the compounds
have the highest Curie temperatures, which do not exceed 400 K however (370 K
for (La0.7 Sr0.3 )MnO3 ; 300 K for (La0.7 Ca0.3 )MnO3 ). These materials are thought
to be half-metallic, and they exhibit the remarkable phenomenon of colossal
magnetoresistance just below the Curie temperature, especially when they are in thin
film form. A huge peak in the resistance, related to the difficulty of electron hopping
among the Mn4+ cores when the spins begin to become disordered, is dramatically
reduced when the Mn spins are aligned by a large applied magnetic field (Fig. 17).
On further doping, toward x = 0.5, the Mn3+ and Mn4+ ions tend to order on
alternate sites, and the oxides become insulators again, adopting C-type antiferro-
magnetic order. Not only is there charge order, but the eg electrons of the Mn3+
3d4 ions undergo orbital order, with correlated Jahn-Teller distortions on nearby
sites. The CE type of charge, orbital, and spin order that may arise near x = 0.5 is
illustrated in Fig. 18. When x = 0.67, other types of stripes of charge and orbital
order can appear, and when x ≈ 1, the antiferromagnetism becomes G-type. The
rich phase diagrams of the mixed-valence manganites reflect the complex balance

Fig. 17 Colossal
5
magnetoresistance of 0T
(La0.7 Ca0.3 )MnO3 in a field 10 T
of 10 T
4
Resistivity (Ω mm)

0
0 50 100 150 200 250 300 350
Tempetature, T (K)
884 J. M. D. Coey

Fig. 18 Type CE order in


(Nd0.5 Sr0.5 )MnO3 , showing
charge and orbital order.
Circles are Mn4+ ; elongated
orbitals are Mn3+

Fig. 19 The structural and La1-xSrxMnO3


magnetic phase diagram of
the mixed-valence
manganites (La1-x Srx )MnO3 R
O T H
[after 38]. Paramagnetic 400
green, ferromagnetic blue,
antiferromagnetic orange;
metallic light color, insulating
dark color; O orthorhombic,
T (K)

O’ sro
R rhombohedral, T R T
tetragonal, M monoclinic, H 200 T H
hexagonal. ‘ charge order, ‘’ M
orbital order. A, C, and G C G
refer to the antiferromagnetic A
O’ sro
structures; C, canted; sro,
c A O’’ O
short range order
0
0.0 0.2 0.4 0.6 0.8 1.0
concentration x

of coulomb, elastic, and magnetic interactions in these materials. The diagram for
(La1-x Srx )MnO3 is shown in Fig. 19 [54].
It must be admitted that the purely ionic picture of a highly polar electronic
structure in these and other 3d oxides is greatly oversimplifies – the Mn-O bonds
are substantially covalent, and the charge transfer from cation to anion is much less
than the formal charge states suggest. Nevertheless, the magnetic properties of the
manganese ions are well described by the quantum numbers S = 3/2 or S = 2 despite
the integrated charge in the vicinity of the cations being closer to 1 or 2, rather than
3 or 4.

Double Perovskites and Related Materials


The double perovskites have the formula A2 BB’O6 . The tetragonal unit cell with
√ √
lattice parameters ∼ 2a0 × 2a0 × 2a0 (Fig. 15b) contains four formula units,
17 Magnetic Oxides and Other Compounds 885

Table 15 Some examples of magnetic double perovskites [87, 107]


Oxide a(pm) b(pm) C(pm) Tc (K) Oxide a(pm) b(pm) c(pm) Tc (K)
Sr2 CrMoO6 557 791 473 Ca2 FeReO6 541 553 769 539
Sr2 CrWO6 784 458 La2 CoMnO6 553 551 780 204
Sr2 CrReO6 552 780 635 Lu2 CoMnO6 516 555 742 48
Sr2 CrOsO6 556 1360 725 La2 NiMnO6 547 551 775 280
Ca2 CrReO6 539 546 766 360 Y2 NiMnO6 523 555 749 85
Ca2 FeMoO6 541 553 770 345 Sr2 CrWO6 786 458
Sr2 FeMoO6 558 790 436 Sr2 FeWO6 565 794 39
Ba2 FeMoO6 570 807 367 Ca2 CrWO6 639 545 766 161
Pb2 FeTiO6 393 404 543 Ca2 NiWO6 540 554 769 56
Sr2 FeReO6 556 780 405 Cd2 FeNbO6 385 388 386 48
Sr2 FeOsO6 786 790 140 La2 CoTiO6 556 557 785 25
Bold, ferromagnetic; normal, ferrimagnetic; italic, antiferromagnetic

but the tolerance factor influences the structure, much as it does for the single
perovskite oxides. There are about a thousand known compounds, and a quarter
of them order magnetically. Most are insulating and antiferromagnetic with Néel
temperatures below 50 K, but tens of them are ferromagnetic or ferrimagnetic at
relatively high temperature, and some of these are half-metallic [87, 107]. Magnetic
ordering temperatures are sensitive to the degree of atomic order on the cation
sites. Perhaps the best-known example is Sr2 FeMoO6 , which is half-metallic with
TC = 436 K. Ferromagnetic exchange is promoted by the Mo, formally in a 4d1
state. More examples are listed in Table 15. Compounds with Re or Os often order
at the highest temperatures, but the spin-orbit coupling of the heavy metal tends to
mix the spin channels and compromise the half-metallicity.
There are very many layered perovskites, composed of slabs of the basic
ABO3 structure, with an intercalated motif. YBa2 Cu3 O6 , the end-member of the
famous series of high-temperature cuprate superconductors, is antiferromagnetic,
with TN = 420 K, and a magnetic moment of 0.5 μB on the Cu2+ sites, where
the 3d electron is in a highly anisotropic x2 - y2 orbital. Residual antiferromagnetic
correlations are thought to play a role in the superconductivity of the compounds
with more oxygen.

Pyrochlores

The pyrochlore structure of A2 B2 O7 compounds has a large cubic unit cell


containing Z = 8 formula units, shown in Fig. 20. The structure is related to that
of fluorite. Both A- and B-sites form lattices of corner-sharing tetrahedra with an
oxygen at the center. Pyrochlore itself is an important niobium mineral, but the
structure is of interest here because of the wide range of interesting magnetic and
electrical properties exhibited by synthetic pyrochlore oxides [45]. A may be a
rare-earth or alkaline earth ion, and B is often a transition-metal ion. The B-site is
886 J. M. D. Coey

Fig. 20 Crystal structures (a) fluorite-type AO2 and (b) pyrochlore A2 B2 O7 ; (c) shows the
connectivity of the A-site tetrahedra, each with an oxygen ion at the center

octahedrally coordinated by oxygen, whereas the A-site is eightfold coordinated, as


in perovskite, but the key feature is that the lattice of corner-sharing metal tetraheda
is fully frustrated whenever the A-A or B-B interactions are antiferromagnetic.
The topology of a frustrated lattice forbids the simultaneous satisfaction of all the
exchange bonds. Other examples are the planar triangular lattice, and the lattice
formed by the B-sites in the spinel structure, but the versatility of the pyrochlore
structure makes it ideal for investigating magnetic frustration in oxides. There exist
a huge number of possible compounds and solid solutions; we give a few illustrative
examples.
The prototype spin ice compounds with A = Dy or Ho and B = Ti have
nonmagnetic Ti4+ on B-sites and a trivalent rare earth with a large moment of
10 μB on A-sites. The exchange coupling among R3+ ions is weak (Table 19)
and smaller than the dipole-dipole interaction [11]. The arrangement of A-site
tetrahedra, each with an oxygen at the center, is shown in Fig. 20c). The oxygen
creates a crystal field along a < 111> direction, which stabilizes an Ising-like ±15/2
or ± 8 doublet with strong axial anisotropy. Two rare-earth moments point inward
and two outward in each tetrahedron. This is analogous to the ice-rule for hydrogen
bonds; an oxygen atom in ice forms two long and two short hydrogen bonds with
its four oxygen neighbors. The magnetic ground state that is reached well below
1 K has substantial frozen-in entropy, which agrees with that predicted by Pauling
for real ice. Hence, these pyrochlores are called dipolar spin ices. A chain of flipped
spins running through the spin ice has opposite magnetic charges at each end, which
behave as separated magnetic monopoles [15]. All spins can be flipped to the nearest
<111> direction at low temperature in a small applied field to give a net moment of
5μB /Dy.
Other pyrochlores (Table 16) with rare-earth ions on A-sites and nonmagnetic
ions on B-sites exhibit complex ferromagnetic (e.g., Tb2 Sn2 O7 ) or antiferromag-
netic (e.g., Gd2 Sn2 O7 , Er2 Ti2 O7 ) order below about 1 K, due to the interplay of
weak exchange and dipolar interactions in the presence of much stronger crystal
field effects on the rare earth. Tb2 Ti2 O7 does not order. It is a spin liquid with
dynamic antiferromagnetic correlations, but no spin freezing or magnetic order
down to 50 mK.
17 Magnetic Oxides and Other Compounds 887

Table 16 Examples of some magnetic oxides with the pyrochlore structure


Compound a0 (pm) Magnetism Tc (K) Ms (kAm−1 )
Dy2 Ti2 O7 1014 Spin ice ∼1 ∼711
Tb2 Ti2 O7 1017 Spin liquid – –
Y2 Mo2 O7 1021 Spin glass 22.5 ∼0
Y2 Ru2 O7 1011 Four-sublattice antiferromagnet 76 0
Tl2 Mn2 O7 990 Half-metallic ferromagnetic semimetal 118 318

The B-sites of the pyrochlore lattice also form a network of corner-sharing


tetrahedra, which is fully frustrated when the nearest-neighbor interactions are
antiferromagnetic. In Y2 Mo2 O7 , the Mo4+ has a 5d2 configuration with S = 1
and strong antiferromagnetic Mo-O-Mo superexchange; the paramagnetic Curie
temperature θp ≈ −200 K. Spin glass freezing occurs at 22.5 K.
In Y2 Ru2 O7 , the Ru4+ has a low-spin 4d4 configuration with S = 1. There is a
transition at 76 K to a four-sublattice antiferromagnetic state. The transition is at a
higher temperature when A is a magnetic rare earth (e.g., 160 K for Pr).
Tl2 Mn2 O7 exhibits colossal magnetoresistance near the ferromagnetic Curie
temperature of 118 K. Its ferromagnetism is due to ferromagnetic Mn4+ -O-Mn4+
superexchange interactions. A small overlap between Tl 6 s and Mn 3d bands makes
the material semimetallic, and the splitting of the 6 s band gives it a half-metallic
character.

4d and 5d Oxides

On passing from 3d to 4d to 5d oxides, the d electrons become progressively more


delocalized, and the Hund’s rule exchange JH and the on-site Coulomb repulsion U
become weaker. On the other hand, the spin-orbit interaction becomes increasingly
important in heavy elements. The 5d and 5f compounds exhibit spin-orbit coupling
that may be as large as JH , so the normal L-S coupling that is the basis of Hund’s
rules breaks down, and it must be replaced by j-j coupling, where spin and orbital
angular momentum are first coupled together for the individual electrons.
A consequence of weakening JH is the appearance of low-spin states when the
cations are subject to a crystal field produced by the coordinating oxygen ions.
Table 17 shows the electronic configurations of some 3d, 4d, and 5d ions, which
may be found in high formal charge states in oxides.
RuO2 is a rutile-structure Pauli paramagnet, but the perovskite ruthenate,
SrRuO3 , is a ferromagnetic 4d band metal with TC = 165 K and a moment of 1.6
μB per formula unit [14]. CaRuO3 has a similar atomic configuration to SrRuO3 ,
but it does not order magnetically. The d4 ion Ru4+ may have a spin-orbit singlet
ground state [67]. Examples of intercalated perovskites are the Ruddlesden-Popper
phases Sr(n + 1) Run O(3n + 1) . The simple perovskite SrRuO3 can be considered as
the n = ∞ member of the series. The n = 1 and n = 2 compounds Sr2 RuO4 and
888 J. M. D. Coey

Table 17 Configurations of Cr,Mo,W Mn,(Tc),Re Fe,Ru,Os Co,Rh,Ir


dn ions, showing n for
3+ 3 4 5 6
different charge states
4+ 2 3 4 5
5+ 1 2 3 4

Sr3 Ru2 O7 are, respectively, a p-wave, spin triplet superconductor with Tsc ≈ 1 K
[68] and a quasi-two-dimensional metal close to a quantum critical point, with a
metamagnetic spin-density wave transition [66].
5d and 5f compounds exhibit the strongest spin-orbit coupling. The spin and orbit
are coupled together for the individual electrons, to form j = l ± ½ states. The 5d
states are extended and tend to form broad bands, with relatively weak correlations.
The crystal field in octahedral sites exceeds the on-site JH that leads to Hund’s rules
in 3d ions. They are therefore in a low-spin state, where the t2g orbitals will be
occupied before the eg orbitals, in octahedral coordination. The t2g manifold has an
effective orbital quantum number l = 1, so the one-electron j-states are j = ½ and
j = 3/2. The possible mj states for the six electrons are a lower j = 3/2 quartet and
an upper j = ½ doublet, as illustrated in Fig. 20. The t2g wave functions are then
mixtures of the xy, yz, and zx states, where the spin-orbit interaction mixes the spin
and orbital character. For example, Ir3+ (t2g 6 ) has fully occupied t2g orbitals, and
it is in a nonmagnetic low-spin state in SrIrO3 , which is a nonmagnetic semimetal,
whereas for Ir4+ (t2g 5 ), the j = 3/2 states are fully occupied, and one electron is in a
j = ½ state, with wave function:

1 1  √ 
| , + = 1/ 3 (| xy, ↑ + | yz, ↓ + i | zx. ↓ )
2 2
1 1  √ 
or | , − = 1/ 3 (− | xy, ↓ + | yz, ↑ − i | zx. ↑ ) (8)
2 2

The Ruddlesden-Popper compound Sr2 IrO4 is a spin-orbit insulator, rather like a


Mott insulator where λ plays the role of U. It exhibits canted antiferromagnetic order
below TN = 240 K. Sr3 Ir2 O7 also has Ir with a 5d5 configuration, and it is a collinear
antiferromagnet with TN = 285 K. The compounds Na2 IrO3 and Li2 IrO3 with Ir
in the same J = ½ configuration order antiferromagnetically at the much lower
temperature of 15 K. The structure of these oxides is a stack of planar honeycomb
sheets, where the J = ½ iridium has a tendency to form a ferromagnetic spin liquid
due to frustration inherent in the wave function (15). This Kitaev magnetism is
thought to compete with the antiferromagnetic order [100].

4f Oxides

The rare earths normally adopt a trivalent ionic configuration, but Eu and Yb
are exceptions, preferring the stable closed (sub)shell configurations 4f7 and 4f14 ,
17 Magnetic Oxides and Other Compounds 889

Table 18 Magnetic properties of europium chalcogenides


Oxide Lattice parameter (nm) Magnetic order Tc (K) J1 (K) J2 (K)
EuO 516 Ferromagnetic 69.3 0.60 0.12
EuS 596 Ferromagnetic 16.6 0.23 −0.11
EuSe 620 FerrimagneticAntiferromagnetic 4.62.8 0.16 −0.16
EuTe 661 Antiferromagnetic 9.6 0.10 −0.21

Table 19 Magnetic ordering R2 O3 Gd Tb Dy Ho Er Yb Lu


temperatures of some R2 O3
a0 (nm) 1081 1072 1067 1060 1055 1049 1039
sesquioxides
TN (K) 1.7 2.4 1.2 2.0 3.4 2.3 –

respectively, that correspond to the divalent ionic configuration. For similar reasons,
Ce tends to adopt a quadrivalent 4f0 configuration.
EuO crystallizes in the cubic NaCl structure (Fig. 11d), with a0 = 516 pm.
It is a ferromagnetic insulator with a bandgap of 1.2 eV and TC = 69 K. The
bandgap decreases by 0.2 eV at TC due to spin splitting of the empty 5d conduction
band. When nonstoichiometric (cation deficient) or doped with a trivalent rare
earth such as Gd, the Curie temperature of EuO is increased substantially by the
presence of 5d conduction electrons. A metal-insulator transition appears at TC ,
where the conductivity can change by up to 14 orders of magnitude; the transition is
accompanied by colossal negative magnetoresistance of up to 108 % per tesla. The
carriers are magnetic polarons [86].
The isostructural europium monochalcogenides (Table 18) exhibit a systematic
variation of magnetic order governed by the relative strengths of the J1 and J2
interactions. Unlike the nearest-neighbor interaction in the 3d monoxides (§ 5.1.1),
J1 is positive, and it competes in the chalcogenides with a negative J2 coupling
cube-corner cations via 180◦ superexchange bonds.
Rare-earth sesquioxides R2 O3 crystallize in the cubic bixbyite structure, with
Z = 16. The magnetic coupling between rare-earth ions is weak, and none of them
order antiferromagnetically above 5 K (Table 19). The hydroxides R(OH)3 have
similarly low Néel temperatures.
There are some 4f mixed-valence compounds. Eu3 O4 is antiferromagnetic below
5.0 K, but it undergoes a metamagnetic transition in a low magnetic field. Eu3 S4 is
cubic at room temperature, but it undergoes a charge ordering transition at 160 K
and orders ferromagnetically at 3.1 K.

5f Oxides and Related Compounds

The actinides in the first half of the series, from Pa to Bk, are sufficiently stable
to allow magnetic measurements to be made on their compounds. The moment
is mainly orbital in character. The elements form a series of dioxides (Table 20)
with the cubic CaF2 fluorite structure shown in Fig. 20a. The actinides also
890 J. M. D. Coey

Table 20 Magnetism of Oxide Ion a0 (nm) Magnetism Tc


actinide dioxides with the
ThO2 5f0 560 Nonmagnetic –
fluorite structure
PaO2 5f1 551 Unknown
UO2 5f2 547 Antiferromagnetic 31
NpO2 5f3 545 Antiferromagnetic 25
PuO2 5f4 540 Nonmagnetic –
AmO2 5f5 538 Antiferromagnetic 9
CmO2 5f6 537 Nonmagnetic –
BkO2 5f7 538 Antiferromagnetic 3

Table 21 Magnetism of US and related NaCl structure actinides [12]


Compound a0 (pm) Magnetism Tc (K) Compound a0 (nm) Magnetism Tc (K)
US 549 Ferromagnet 177 AmN 499 Paramagnet –
USe 575 Ferromagnet 175 CmN 503 Ferromagnet 125
UTe 616 Ferromagnet 104 UP 559 Antiferromagnet 125
UN 489 Antiferromagnet 52 UAs 577 Antiferromagnet 127
NpN 490 Ferromagnet 87 USb 619 Antiferromagnet 241
PuN 490 Antiferromagnet 19

form monoxides and chalcogenides, as well as pnictides, all with the cubic NaCl
structure (Table 21). US has the largest cubic anisotropy of any known material,
K1 c = 43 MJm−3 . The spin-orbit coupling is very strong, and the 5f one-electron
states have l = 3 and s = ±1/2, giving j = 5/2 or 7/2. The j = 5/2 states are split in
the cubic crystal field into a lower doublet and an upper singlet, so that occupancy
by one, two, three, five, or seven electrons leads to atomic moments that can order
magnetically, whereas the atoms with four and six electrons are nonmagnetic [85].
The data summarized in Tables 20 and 21 bear this out.
Cubic actinide compounds often exhibit 3k antiferromagnetic structures, charac-
terized by propagation vectors with Fourier components along all three equivalent
cubic directions. The spins are then aligned along the four <111> directions.

Related Compounds

There are many other families of 3d transition-metal compounds that are more or
less ionic than the oxides. It is useful to review a few of them, so as to place the
oxide magnetism in a broader context.

Halides

The most electronegative anion, F− , forms the strongest ionic bonds with transition
metals. The 3d difluorides have the tetragonal rutile (TiO2 ) structure, illustrated
in Fig. 18b). The difluorides listed in Table 22 are all antiferromagnets, with
17 Magnetic Oxides and Other Compounds 891

Table 22 Properties of difluorides and trifluorides


a (pm) c (pm) TN (K) ac (pm) TN (K)
MnF2 487 331 67 CrF3 354 80
FeF2 470 331 78 MnF3 a 45
CoF2 470 318 38 FeF3 373 363
NiF2 465 308 78 CoF3 365 460
a Monoclinic

relatively low Néel temperatures in the range 38–78 K. The trifluorides crystallize
in a rhombohedrally distorted ReO3 structure, which resembles the ideal perovskite
structure of Fig. 15a) without the A-site cation in the center of the unit cell. The
lattice parameter ac in the table pertains to a cubic ReO3 -type unit cell. The Néel
temperatures with almost straight, unfrustrated M-O-M bonds range up to 363 K, for
FeF3 , indicating weaker superexchange than in comparable oxides such as YFeO3 .
As in the orthoferrites, the symmetry permits a Dzyaloshinskii-Moriya interaction
and canting of the antiferromagnetic sublattices in trifluorides when the moments
lie perpendicular to rhombohedral c-axis [73]. Ferric fluoride has the largest low-
temperature iron sublattice moment and the largest 57 Fe hyperfine field, 62.2 T, of
any ferric iron due to the strongly ionic character of the bonding. The comparable
value for aFe2 O3 is 54.2 T. [52]. FeF3 can also crystallize in a highly frustrated
pyrochlore-like structure, where the Néel temperature is only 22 K.
Tetragonal K2 NiF4 is an insulator with Ni2+ , S = 1, and a model two-
dimensional magnetic material because the interlayer exchange interactions cancel.
Rb2 CoF4 is an almost-ideal two-dimensional Ising antiferromagnetic, thanks to the
strong uniaxial anisotropy experienced by Co2+ .
Unlike the fluorides, which have three-dimension crystal structures similar to
rutile or perovskite, the other transition-metal halides have hexagonal layered
structures [71] because of the large size of the anions. The ionic volumes of Cl− ,
B− , and I− are three to five times larger than that of F− . The crystal structure type is
either CdI2 (Fig. 11i) or CdCl2 (Fig. 21d). The hexagonal layers are the same in each
case, but the packing is ABABAB . . . . in the CdI2 case and ABCABC . . . . . . in
CdCl2 . All order antiferromagnetically in collinear, triangular, or helical structures,
mostly at temperatures of order 10 K although |θp | is an order of magnitude greater
for the Ti and V compounds. The in-plane exchange interactions are competing or
frustrated. Magnetic order in monolayers of the vanadium compounds may set in at
higher temperatures than in the bulk [112] (Table 23).
Trihalides form R3̄ layer structures with a honeycomb lattice that can be readily
exfoliated. The layers are ferromagnetic in CrX3 with X = Cl, Br and I. CrCl3
is a planar antiferromagnet with TN = 14 K and easy-plane anisotropy, whereas
CrBr3 and CrI3 are ferromagnetic with strong easy c-axis anisotropy and TC = 33
K and 61 K respectively. Exfoliated monolayers of CrI3 have Tc = 45 K. FeCl3
is helimagnetic below 9 K, and magnetically-ordered layers can be intercalated
into graphite. The rare earths form trifluorides with very low Curie temperatures.
GdF3 , with TC = 1.2 K, for example, is an excellent potential low-temperature
magnetocaloric material.
892 J. M. D. Coey

Fig. 21 The crystal structures of (a) FeS2 (pyrite), (b) FeF2 (rutile),)(c) FeF3 (ReO3 -type), and
(d) FeCl2 (CdCl2 )

Table 23 Magnetic properties of 3d dihalides


Type a(nm) ca (nm) TN (K) Type a(nm) ca (nm) TN (K)
TiCl2 CdI2 356 588 85 FeCl2 CdCl2 360 583 24 (f ||c)
TiB2 CdI2 363 649 FeB2 CdI2 378 623 14 (f ||c)
TiI2 CdI2 411 682 FeI2 CdI2 403 675 9 (⊥c)
VCl2 CdI2 360 583 36 (120◦ ) CoCl2 CdCl2 354 581 25 (f ⊥c)
VB2 CdI2 377 618 36 (120◦ ) CoB2 CdI2 369 612 19 (f ⊥c)
VI2 CdI2 406 676 15 CoI2 CdI2 396 665 11 (helix)
MnCl2 CdCl2 373 586 2 (helix) NiCl2 CdCl2 348 580 52 (f ⊥c)
MnB2 CdI2 389 627 2 (⊥ c) NiB2 CdCl2 370 609 52 (f ⊥c)
MnI2 CdI2 414 682 3 (helix) NiI2 CdCl2 390 654 75 (helix)
a One-half or one-third of the hexagonal c-axis

Chalcogenides

The elements in column 16 of the periodic table, S, Se, and Te, known as chalcogens
form increasingly covalent bonds in their transition-metal chalcogenides. Some
crystal structures are common to oxides and chalcogenides, but others, such as the
pyrite (FeS2 ) or nickeline (NiAs) structures, never form with oxides. Transition-
metal sulfides exhibit a wide variety of magnetic and electrical properties [84].
The main iron-sulfur minerals are pyrite and pyrrhotite. Pyrite, FeS2, also known
as fool’s gold, is often found as large brassy cubic crystals. The sulfur is present
in the structure as disulfide S2 2− pairs, and the ferrous iron, with formal charge
17 Magnetic Oxides and Other Compounds 893

Fig. 22 Structures of transition-metal sulfides. (a) FeS2 (pyrite), (b) NiS (NaAs-structure),
(c) Fe7 S8 (pyrrhotite), and (d) CuFeS2 (chalcopyrite)

Table 24 Magnetic properties of pyrites


a0 (pm) Spin Magnetism Tc (K)
MnS2 Hauerite Semiconductor 610 5/2 Antiferromagnet 48
FeS2 Pyrite Semiconductor 542 0 Temperature-independent –
paramagnet
CoS2 Cattierite Metal 553 (½) Ferromagnet 120
NiS2 Vaesite Semiconductor 568 1 Antiferromagnet 50
CuS2 Chalcocite Metal 571 ½ Pauli paramagnet –
ZnS2 – Semiconductor 595 0 Diamagnet –

2+, is in an S = 0 low-spin state with no magnetic moment. FeSe2 and FeTe2 are
similarly nonmagnetic. CoS2 is an itinerant ferromagnet with Tc = 120 K. Other
pyrites are antiferromagnetic. Properties are collected in Table 22, and the structure
is illustrated in Fig. 22a) (Table 24).
Chalcopyrite, CuFeS2 , is an important copper ore mineral. It is a tetragonal
antiferromagnetic semiconductor with a = 529 pm and c = 1042 pm; the structure
(Fig. 22d) is related to zinc blende (ZnS). Sulfur is unpaired S2− . The Néel
temperature of 815 K is exceptionally high, and the Fe3+ moment is reduced to 3.85
μB by covalent mixing with sulfur. Bornite, Cu5 FeS4 , another copper-iron sulfide
mineral, is antiferromagnetic below 76 K. Cubanite CuFe2 S3 is an orthorhombic
mixed-valence iron mineral, a canted antiferromagnet with a weak ferromagnetic
moment of just 1 Am2 kg−1 , which transforms to a cubic structure at 490 K.
An important group of magnetic chalcogenides, listed in Table 25, crystallize in
a variant of the hexagonal NiAs (nickeline) structure (a = 344 pm, c = 571 pm;
Fig. 21b)). The compounds are usually substoichiometric, with ordered or disor-
dered cation vacancies. Cr2 S3 , with ordered Cr vacancies, is ferrimagnetic below
130 K. CrSe is a noncollinear antiferromagnet, but CrTe is a ferromagnet. Both of
894 J. M. D. Coey

Table 25 Magnetic properties of hexagonal chalcogenides


Tc (K) Ms (kAm−1 ) K1 (MJm−3 )
Cr2 S3 Ferrimagnet 130 –
CrSe Noncollinear antiferromagnet 279
CrTe Ferromagnet 340 ∼200 –
FeS Troilite Antiferromagnet 588 630 0.01
Fe7 S8 Pyrrhotite Ferrimagnet 598 150 0.32
Fes S4 Smythite Ferrimagnet 600 580 1.20
Fe7 Se8 Achavalite Ferrimagnet 448 70 0.25
Fes Se4 Ferrimagnet 314 ∼50 –
NiS Antiferromagnet 260a –
a First-order transition

them order magnetically close to room temperature. Thin films of CrTe grown on
SrTiO3 substrates (111) show evidence of a topological Hall effect [118].
Pyrrhotite, Fe7 S8 , is a rock-forming mineral and the second most common
strongly magnetic mineral, after magnetite. The iron vacancies in the monoclinic
structure are ordered on alternate planes (Fig. 22c), giving a ferrimagnetic structure,
with moments along the c-axis, and substantial uniaxial anisotropy.
Antiferromagnetic FeS (troilite) has no cation vacancies, but the Fe forms
triangles in a variant of the NiAs structure, stable below a transition at Tα = 413 K.
Hexagonal NiS is a planar antiferromagnet, with a first-order transition to a Pauli
paramagnetic state at 260 K. A trigonal polymorph, Millerite, is a nonmagnetic
metal.
MnS exists in three polymorphic forms; all are antiferromagnetic, with 12
nearest-neighbor manganese atoms [34]. The Mn is octahedrally coordinated by
sulfur in alabandite, αMnS, which has the cubic halite (NaCl) structure. In βMnS
and γMnS, the Mn is tetrahedrally coordinated by sulfur, and the sulfides have the
cubic zinc blende structure or the hexagonal wurtzite structure, respectively. Néel
temperatures are 154 K, 100 K, and 80 K, respectively.
An important family of cubic magnetic chalcogenides are the synthetic sulphos-
pinels that were intensively studied in the 1970s. Most of them have a normal
cation distribution with Cr3+ or Rh3+ on B-sites, which have a strong preference
for d3 ions. The structure was illustrated in Fig. 11c). Natural sulphospinels are
rare and include carrollite, CuCo2 S4 , greigite Fe3 S4 which is the sulfide analogue
of magnetite that is produced by some magnetotactic bacteria, and nonmagnetic
linneite Co3 S4 . Of these, only greigite orders magnetically, at 580 K. A large
number of synthetic compounds have been prepared [105]. The structural and
magnetic properties of the sulphospinel family are summarized in Table 26.
Note that the Rh3+ and the Co3+ in B-sites are in a low-spin t2g state with S = 0,
so any magnetic order depends on the A-site cations. As with the oxide spinels,
there are numerous interesting solid solutions between the various end-members.
One example is Fe0.5 Rh0.5 Rh2 S4 , which orders antiferromagnetically at 145 K,
despite Fe3+ , the only magnetic ions, being separated by 985 pm. Another example
17 Magnetic Oxides and Other Compounds 895

Table 26 Sulphospinels and a0 (pm) Structure Tc (K)


related chalcogenides
MnCr2 S4 1011 Ferrimagnet 70
FeCr2 S4 999 Ferrimagnet Semi 189
CoCr2 S4 923 Ferrimagnet 225
CuCr2 S4 982 Ferromagnet p semi 420
CuCr2 Se4 1033 Ferromagnet p semi 460
CuCr2 Te4 1112 Ferromagnet p semi 365
ZnCr2 S4 998 Antiferromagnet 18
ZnCr2 Se4 1044 Antiferromagnet 20
CdCr2 S4 1024 Ferromagnet Semi 85
CdCr2 Se4 1074 Ferromagnet Semi 130
HgCr2 S4 1021 Antiferromagnet 36
HgCr2 Se4 1075 Ferromagnet p semi 106
Fe3 S4 988 Ferrimagnet 580
Co3 S4 940 Paramagnet
NiCo2 S4 940 Paramagnet
CuCo2 S4 946 Paramagnet
CoRh2 S4 972 Antiferromagnet Semi 400
FeRh2 S4 986 Antiferromagnet 190
Bold, ferromagnetic; normal, ferrimagnetic; italic, antiferro-
magnetic

is CoRh2 S4 , which has Co2+ on A-sites where the superexchange path involves
two anions (or low-spin Rh3+ ), yet the Néel temperature is 400 K. Superexchange
interactions can be remarkably long-range in a covalent matrix.
Notable among the sulphospinels are the materials which, like EuO, are ferro-
magnetic semiconductors. The Curie temperatures of p-type CuCr2 S4 and CuCr2 Se4
are well above room temperature. A series of chalcogenide layer compounds
including CrSi2 Te6 and Cr2 Ge2 Te6 with van der Waals interlayer bonding can be
exfoliated, like CrI3 , to yield ferromagnetic monolayers. Their Curie Temperature
is ≈ 60 K [48].

Pnictides

The elements in column 15 of the periodic table, N, P, As, Sb, and Bi, are known
as pnictogens and form pnictides with transition elements. The bonding tends to be
metallic rather than ionic, but we consider them briefly here. The iron nitrides are an
important group of magnetic compounds. Nitrogen is one of the few elements in the
periodic table that is small enough to fit into the interstices of a metallic 3d lattice.
Since exchange, especially in iron metal, is highly sensitive to interatomic spacing,
there can be big changes in magnetism whenever nitrogen enters the structure. A
nice example is Fe4 N, where the nitrogen occupies the cube center position and
expands the fcc structure of γ-iron. Fcc iron is at the limit for the appearance of
896 J. M. D. Coey

Table 27 Lattice parameters and magnetism of iron nitrides


a (nm) b (nm) c (nm) TC (K) Ms (kAm−1 )
αFe97 N3 287 1010 1710
α Fe90 N10 283 312 835
αFe16 N2 572 629 810 1830
γ Fe4 N 380 769 1510
εFe3 N 270 436 567
ζFe2 N 483 553 443 9 –

Table 28 Magnetic properties of Mn pnictides


Ms (kAm−1 )
a (pm) b (pm) c (pm) Tc (K) at 300 K
MnN 426 419 Antiferromagnet 660 – Distorted NaCl
structure
MnP 524 318 590 Ferromagnet 291 ∼0 Helix → ferromagnetic
transition at 50 K
MnAs 375 570 Ferromagnet 318 630 First-order NiAS →
MnP structural
transition at TC
MnSb 414 579 Ferromagnet 573 770
MnBi 428 611 Ferromagnet 633 576

magnetism, but Fe4 N is a strong ferromagnet, with TC = 769 K. The cube corner 1a
site and face center 3c sites are now inequivalent, and 1a → 3c charge transfer leads
to a buildup of moment on 1a sites (3.0 μB ) at the expense of 3c sites (2.0 μB ).
The magnetization Ms is still less than it is for αFe (a0 = 286.6 pm), because of
the lattice expansion caused by interstitial nitrogen. The properties of this and other
Fe-N phases are summarized in Table 27 [25].
Tetragonal α”Fe16 N2 is a remarkable iron compound, where the magnetization
actually exceeds that of αFe. The uniaxial anisotropy is approximately 1 MJm−3 , a
value that is much too large for a soft magnet but insufficient for a dense permanent
magnet that would not have to rely, like alnico, on shape anisotropy to retain a
permanently magnetized state [95].
An interesting series of Mn pnictides, Table 28, includes MnAs, MnSb, and
MnBi, which all crystallize in the NiAs structure (Fig. 22b). MnN and MnP
crystallize in an orthorhombic variant. MnBi exhibits uniaxial anisotropy K1 = 0.9
MJm−3 and good magneto-optic properties due to strong spin-orbit coupling on Bi.
In this case, the anisotropy is just sufficient to make a permanent magnet [95].
The rare-earth nitrides RN are ferromagnetic, with a maximum Curie temperature
Tc = 70 K for GdN. All the others have lower Curie temperatures varying as the de
Gennes factor of the rare-earth ion.
17 Magnetic Oxides and Other Compounds 897

Silicates and Carbonates

Silicates

The aluminosilicate minerals, which make up the bulk of the Earth’s crust, are
traditionally classified in terms of the connectivity of the SiO4 tetrahedral units
that are the common feature of all silicate structures. According to the mineralogists
clasification, the tetrahedral units may form a three-dimensional network in tectosil-
icates such as quartz, two-dimensional layers in phyllosilicates such as mica, double
or single one-dimensional chains in the inosilicates such as amphibole or pyroxene,
rings in cyclosilicates such as tourmaline, and isolated single or double tetrahedral
groups in silicon-poor nesosilicates such as olivine or sorosilicates such as epidote.
The other structural elements in aluminosilicates – octahedrally coordinated metals
(aluminum, iron) and alkali or alkaline earth cations in larger sites – form networks
that are interspersed with those of the silica tetrahedra [58].
Here we focus on the iron-rich silicates that are end-members of the rock-forming
solid solution series. A classification based on the connectivity of the octahedral
units, which often share edges or corners, is more appropriate for discussing the
magnetism [23]. The illustrations of structures in this section therefore show the
metal cation connectivity, as well as the silica frameworks (blue tetrahedra in
Fig. 23). Edge-sharing octahedra give rise to weak ferromagnetic superexchange,
but no silicate is known to order ferromagnetically, and none of them order
antiferromagnetically above 160 K. The relatively low density of magnetic ions
and the reduced dimensionality of the basic magnetic sheets, ribbons, or chains all
contribute to the low magnetic ordering temperatures.
Magnetic order in the iron end-members of the main structural families will
now be reviewed, and a summary is provided in Table 29. Olivines, with isolated
SiO4 tetrahedra, are framework structures from a magnetic viewpoint. In the
orthorhombic structure, space group Pnma, there are two octahedral iron sites,
M1 chains along b and M2 sheets in the ac plane (Fig. 23a). The orthorhombic
ferrous olivine fayalite Fe2 SiO4 orders antiferromagnetically at 65 K. The magnetic
structure is one of ferromagnetic M1 chains with moments slightly canted with
respect to b, coupled antiparallel to each other, and to the M2 neighbors. The

Fig. 23 Silicate crystal structures. (a) Olivine, (b) trioctahedral 1:1, (c) dioctahedral 2:1, with
intercalated ions, and (d) amphibole
898 J. M. D. Coey

Table 29 Magnetic properties of some iron-rich silicates [23]


Cation Magnetic
Mineral Ideal formula structure TN (K) θ p (K) structure
Framework structures
Fayalite {Fe2 2+ }[Si]O4 M1 chains 65 −37 Canted AF
M2 sheets chains
Laihunite {Fe2+ Fe2 3+ }[Si2 ]O8 M1 half 160 – Crazy spins
empty AF planes
M2 planes
Sheet structures
Greenalite {Fe3 2+ }[Si2 ]O5 (OH)4 1:1 sheets 17 24 Metamagnetic
Cronstedtite {Fe2 2+ Fe3+ }[SiFe3+ ]O5 (OH)4 1:1 sheets 10 −16 AF sheets
Minnesotaite {Fe3 2+ }[Si4 ]O10 (OH)2 2:1 sheets 28 38 Metamagnetic
Feripyrophylite {Fe2 3+ }[Si4 ]O10 (OH)2 2:1 sheets 18 −22 AF sheets
Chain structures
Hedenbergite
ca {Fe2+ }[Si2 ]O6 M1 chains 38 35 AF, F chains
Acmite
Na {Fe3+ }[Si2 ]O6 M1 chains 8 −46 AF chains
Orthoferrosilite {Fe2+ 2 }[Si2 ]O6 M1.M2 41 28 AF, F
ribbons ribbons
Crocidolite
Ca2 {Fe3 2+ Fe2 3+ }[Si8 ]O22 (OH)2 M1,M2,M3 30 27 AF, F
ribbons ribbons
Grunerite {Fe7 2+ }[Si8 ]O22 (OH)2 M1,M2,M3, 47 67 AF, F
M4 ribbons ribbons
Ilvaite
ca {Fe2 2+ Fe3+ }[Si2 ]O8 (OH)2 Double A 116 −300 AF ribbons
ribbons
Group structures
Almandine
Fe3 2+ {Al2 }[Si3 ]O12 Isolated 7 – AF
24c
Andradite
Ca3 {Fe2 3+ }[Si3 ]O12 Isolated 11 – Canted AF
16a

crystal field acting on the Fe2+ favors different directions at the two sites, leading
to alternate spin canting at both sites.
Laihunite is an iron-deficient ferric olivine where half the M1 sites are vacant,
and the ferric iron in the M2 planes orders antiferromagnetically at 160 K. The
remaining ferrous ions in the M1 chains are crazy spins that are initially decoupled
from the magnetic order because they have equal numbers of neighbors on the two
M2 antiferromagnetic sublattices and cannot decide which way to turn.
Natural olivines are generally solid solutions; the forsterite (Mg2 SiO4 )-fayalite
series is a good example. Synthetic olivines with Mn, Co, or Ni also have canted
antiferromagnetic structures, but with lower Néel temperatures than fayalite [44]
(Table 29).
Sheet silicates are layered materials from the point of view of both the silicon
and the metal cations. The family includes rock-forming minerals like mica and
17 Magnetic Oxides and Other Compounds 899

clay minerals with a particle size below 2 μm that are plentiful in sediments and
soils. There are two main families, the 1:1 minerals, with an octahedral/tetrahedral
bilayer, and the 2:1 minerals, with a tetrahedral/ octahedral/tetrahedral sandwich
trilayer (Fig. 23b, c). The octahedral sites form a triangular lattice and are all
occupied by divalent cations in the trioctahedral minerals, but only two-thirds of
them are occupied by trivalent cations in the dioctahedral minerals, either at random
or forming a honeycomb lattice. Various stacking sequences of the layers with
different c parameters are possible, and interlayer van der Waals interactions are
weak, so it is possible to intercalate water, large monovalent cations, or various
organic molecules in the interlayer space. These silicates form a very interesting
group of two-dimensional materials, not least because of their vital role in soils.
We look first at the divalent, ferrous trioctahedral materials. Greenalite
{Fe2+ 3 }[Si2 ]O5 (OH)4 , the ferrous mineral related to serpentine, is a 1:1 mineral that
orders at 17 K as a planar antiferromagnet with in-plane ferromagnetic octahedral
sheets very weakly coupled by long-range interlayer antiferromagnetic interactions.
The mineral exhibits ferromagnetic hysteresis below TN , due to a spin flop. The 2:1
counterpart is minnesotaite {Fe2+ 3 }[Si4 ]O10 (OH)2 , which behaves similarly. It is
a planar antiferromagnet with TN = 27 K. A neutron diffraction pattern showing
antiferromagnetic reflections is shown in Fig. 24. The interplane exchange is very
weak, but the intraplane exchange is ferromagnetic, with θp = 17 K. The in-plane
magnetization in the octahedral sheets is explained by a trigonal crystal field acting
on the Fe2+ , which splits the 5 T2g triplet giving a 5 A1g ground state in the cubic
crystal field that is split further by a trigonal component to give a lowest Sz = 0
singlet (Fig. 25). This singlet implies easy-plane magnetic anisotropy in the plane
perpendicular to Oz. Biotite, a more chemically disordered 2:1 silicate that often
appears with magnetite inclusions, exhibits spin freezing only below 7 K.
Cronstedite {Fe2+ 2 Fe3+ }[SiFe3+ ]O5 (OH)4 is an unusual mixed-valence triocta-
hedral 1:1 layer mineral, which exhibits the highest conductivity of any silicate. It
orders antiferromagnetically at 12 K, although the paramagnetic Curie temperature
θp is positive, as it is for all the other ferrous trioctahedral silicates, because of the
ferromagnetic Fe2+ -O-Fe2+ superexchange.
Turning now to the ferric dioctahedral sheet silicates, pyrophyllite, with an
ordered honeycomb occupancy of two-thirds of the octahedral sites, orders anti-
ferromagnetically below 18 K, whereas nontronite, which has a similar value of
θp = −20 K but a more disordered cation distribution with a proportion of triangles,
shows spin freezing only below 2 K. Frustration can become an issue in the
trioctahedral minerals with random octahedral site occupancy (Fig. 26).
Next, we come to a large group of silicates with ribbons or chains of metal
cations, including the amphibole and pyroxene groups. There are resemblances
with the 2:1 sheet silicates, where the ribbons can be regarded as strips of a 2:1
layer in a base-centered lattice, with silica tetrahedra between them; the inter-
ribbon exchange coupling should be much weaker than the intra-ribbon exchange,
so ribbons can be regarded as isolated magnetic motifs. In amphiboles, they are four
octahedra wide, whereas in pyroxenes they are only two octahedra wide. The ferrous
900 J. M. D. Coey

50

σ(Am2kg–1)
00 1
Counts

2 25

1
I(T)/I(0)

2 1 1 2
μ0H (T)
0
0 10 20 30
T(K) –25
00 3 00 5 00 7
2 2 2

0 –50
0 10 20 30 40
q(deg)

Fig. 24 (a) Magnetic neutron diffraction pattern of minnesotaite showing the l/2 c-axis reflections
from the planar antiferromagnetic order. (b) A magnetic hysteresis loop measured at 4.2 K
exhibiting a spin flip in about 0.5 T [3]

Fig. 25 Crystal field model


for Fe2+ in sheet silicates

amphibole grunerite has iron in all four sites. θp is positive, 67 K, but grunerite
nevertheless orders antiferromagnetically below the Néel temperature of 47 K with
a collinear structure where the ferromagnetic moments of each of the ribbons that
are magnetically aligned along the b axis order antiparallel to the moments of the
four neighboring ribbons. Crocidolite (blue asbestos) is a fibrous alkali amphibole
having Na in the outer M1 sites of the ribbon, charge compensated by Fe3+ in M2
sites. It orders antiferromagnetically at 30 K in a similar way to grunerite.
The ribbons in pyroxenes are two octahedra wide. Both sites are occupied
by ferrous iron in orthorhombic orthoferrosilite {Fe2+ 2 }[Si2 ]O6 . The exchange is
predominantly ferromagnetic with θp = 27 K, but the magnetic structure below the
17 Magnetic Oxides and Other Compounds 901

Fig. 26 Metal site occupancy in silicates: (a) trioctahedral sheet and (b) ordered and (c)
disordered dioctahedral sheets, (d) amphibole, (e) pyroxene, and (f) ilvaite

ordering temperature of 40 K is again one of ferromagnetic chains with moments


along b, coupled antiferromagnetically to their four neighbors. Hedenbergite and
acmite are, respectively, calcium and sodium clinopyroxenes,
Ca {Fe2+ }[Si2 ]O6
and
Na {Fe+ }[Si2 ]O6 , with predominantly positive and negative superexchange
interactions, respectively. Their Néel temperatures are 38 K and 8 K.
Ilvaite,
Ca {Fe2+ 2 Fe3+ }[Si2 ]O8 (OH)2 , is a mixed-valence silicate with a differ-
ent structure. It orders antiferromagnetically at 116 K, with some frustration.
The last class of magnetically ordered silicates has isolated groups of metal
cations. These include the silicate garnets almandine
Fe2+ 3 {Al2 }[Si3 ]O12 and
andradite
Ca3 {Fe3+ 2 }[Si3 ]O12 having ferrous ions on isolated dodecahedral sites
and ferric ions on isolated octahedral sites, respectively. Nevertheless, they manage
to order antiferromagnetically at 4.2 K and 12 K, respectively, demonstrating that
second-neighbor superexchange is operative here, as it is between the ribbons in
pyroxenes and amphiboles.
In summary, the main reason for the weakness of the principal exchange
interactions in magnetic end-member silicates is edge-sharing octahedra in their
structures. There is little evidence for unusual critical behavior associated with their
low dimensionality because the ratio of exchange interactions within and between
the structural units is only about 10. Weak antiferromagnetic interactions propagate
via the SiO4 tetrahedra.
The ferrous sheet and ribbon silicates all exhibit anisotropic susceptibility above
their Néel temperature, due to the trigonal crystal field acting on the ferrous ions.
The example of data on fayalite is shown in Fig. 27.
Natural iron silicates are generally solid solutions with nonmagnetic cations.
In aluminosilicates, the iron is often a minor component in a solid solution,
replacing aluminum, if trivalent, or an alkaline earth cation if divalent. Iron-rich
end-members are relatively uncommon. When the iron content in a solid solution
is low, superexchange percolation among magnetic cations separated by an oxygen
anion is truncated, and long-range magnetic order becomes impossible, even at low
temperatures. Percolation is possible when the iron concentration on the appropriate
sites exceeds the percolation threshold xp , and at higher magnetic concentrations,
a linear increase of Néel temperature with x up to the iron end-member value is
902 J. M. D. Coey

Fig. 27 (a) Anisotropic


susceptibility of a fayalite
crystal [4]

Table 30 Magnetic properties of olivines [44]


Lattice param-
eters a,b,c TN M0 M1 M0 M2 Magnetic
(pm) Pnma (K) θ (K) (μB ) (μB ) structure
Mn2 SiO4 Tephroite 1060; 626; 490 47 −170 a,b,c 3.85 4.68 cAF M1
chains S||a
Fe2 SiO4 Fayalite 1047; 609; 482 65 −107 a,b; 4.40 4.41 cF M1
−66 c chains S||b
Co2 SiO4 – 1030; 600; 478 50 −102 a; 3.90 3.64 cF M1
−23 b; chains S||b
−59 c
Ni2 SiO4 Liebenbergite 1012; 591; 473 34 – – F M1
chains

Table 31 Magnetic properties of carbonates, phosphates


Mineral Ideal formula Space group Lattice parameters (pm) TC , TN (K) Order
Siderite FeCO3 R3c 469; 1538 38 AF
Rhodochrosite MnCO3 R3c 478; 1567 32 wF
Vivianite Fe3 (PO4 )2 8H2 O B2/m 1009; 1344; 470; 104.3◦ 400 AF

normally to be expected (Fig. 2a). A common feature of the iron-rich phases is that
nonmagnetic cations are generally present on some of the sites, leading to broadened
magnetic ordering transitions. The forsterite-fayalaite (Mg1-x Fex )2 SiO4 series is a
good example. There is no magnetic order below the percolation threshold xp , but
with increasing x, the Néel temperature increases linearly up to the fayalite value of
65 K. (Tables 30 and 31).
17 Magnetic Oxides and Other Compounds 903

Carbonates, Phosphates

Carbonates, like silicates, may order magnetically at low temperature. Two exam-
ples, each with the calcite structure, are siderite FeCO3 and rhodochrosite MnCO3
with Néel temperatures of 38 and 24 K, respectively. The manganese carbonate
was one of the first antiferromagnets where spin canting due to the Dzyaloshinskii-
Moriya interaction was identified, and there is a very sharp susceptibility peak at
the Néel temperature [40, 73]. The ferrous carbonate, vivianite Fe3 (PO4 )2 .8H2 O,
orders in a noncollinear antiferromagnetic structure at 9 K. The different signatures
of the three minerals in their low-temperature susceptibility may allow them to be
quantified in small concentrations in sedimentary mixtures [43].

Oxide Thin Films

The ability to deposit large-area oxide and metal thin films that are smooth to
within atomic dimensions has transformed research in magnetism and made spin
electronics possible [113]. Here we discuss magnetism of oxide thin films, bilayers,
multilayers, and superlattices, especially surface and interface effects. Nonmagnetic
oxide films play essential roles as tunnel barriers, cap, and buffer layers. Figure 28
illustrates these different structures.

Substrates, Caps, and Buffers

Epitaxial magnetic oxide thin films are best grown on nonmagnetic oxide crystal
substrates with similar lattice parameters. The crystal surface frequently presents

Fig. 28 Oxide thin film structures on a substrate. (a) Capped film, (b) pinned tunnel junction,
(c) bilayer, (d) multilayer
904 J. M. D. Coey

square, cubic, or hexagonal oxygen packing, and epitaxial growth usually requires
a lattice mismatch of less than 2%. Strain is relaxed over a length scale of order
100 nm in the film thickness. Certain properties of magnetic thin films, such as
their crystal and electronic structure, magnetic structure, and magnetocrystalline
anisotropy, may be very sensitive to substrate-induced strain, which opens many
possibilities for strain engineering. For example, a film of cubic ferromagnet may
become tetragonal and acquire perpendicular magnetic anisotropy when deposited
on a mismatched substrate where it is subject to biaxial strain. There is a wide
choice of oxide substrates, including functional ones – conducting, ferroelectric, or
piezoelectric – which can be used to tune the magnetism of the thin film. A list of
substrates and their magnetic susceptibility is given in Table 32.
Since the thickness of the substrate, usually 0.3–1.0 mm, is much greater than
the thickness of the oxide film grown on it, which is typically tens or hundreds of
nm, it is particularly important to be aware of the influence of magnetic impurities
in the substrate, as well as interface defects. MgO, for example, often includes
Fe2+ impurities that produce a strong Curie-law upturn in the susceptibility at low
temperature. Small magnetic moments of order 10−8 Am2 that saturate in an applied
field of approximately 100 kAm−1 frequently appear in 5 × 5 mm2 oxide substrates
at room temperature [60]; they may be due to ferromagnetic impurities introduced
in the polishing process or else somehow associated with oxygen vacancies at the
surface [31].
Epitaxy is less important for oxide cap layers that serve to protect a thin film layer
or heterostructure. Buffer layers can be used to improve the lattice match between
a substrate and the thin film to be grown on it. Since silicon wafers are available in
much larger diameters than single-crystal oxides such as SrTiO3 , it is interesting for
oxide electronics to find buffer layers that will enable large-area oxide thin films to
be grown on silicon. CeO2 is a suitable candidate. Buffer layers may also serve as
diffusion barriers.

Thin Film Preparation

The technology to fabricate device-quality oxide thin films advanced rapidly after
the discovery of the cuprate high-temperature superconductors in 1986. The boom
in high Tsc research was facilitated by the development of the pulsed laser deposition
(PLD) method to prepare laboratory-scale samples of oxide thin films by ablating an
oxide target with repeated femtosecond pulses of UV laser radiation [37]. A further
spur to develop wafer-scale oxide thin films with sub-nanometer scale surface
roughness was the need for reliable and reproducible tunnel barriers in sensors and
memory elements based on tunnel magnetoresistance (Fig. 28b). A brief account of
the main methods for making oxide thin films follows.
Oxide thin films can be prepared by physical or chemical methods. Physical
deposition methods involve a source of material separated by a distance d from the
substrate, which is often heated in the range 400–1000 ◦ C to facilitate film growth
[98]. Several variants are indicated in Fig. 29. At low pressure, the atomic species
Table 32 Single-crystal substrates for oxide thin films
OxideAbbreviation Plane Space group Lattice parameter (pm) Susceptibility10−6 Comments
MgO 100 Fm3m 421 −18 Includes paramagnetic Fe2+
110 impurities
111
NaCl 100 Fm3m 564 −14 Soluble in water
Al2 O3 001 R3c 476: 1300 −18
Sapphire 100
SrTiO3 STO 100 Pm3m 391 −7 ε = 300
SrTiO3 :Nb 100 Electrical conductor
LaAlO3 LAO 100 Pm3m 379 −18
(La,Sr)(Al,ta)O3 100 Pm3m 387 −17 Stable
17 Magnetic Oxides and Other Compounds

LSAT 111
MgAl2 O4 MAO 100 Fd3m 808 −15
BaTiO3 BTO 100 P4mm 399, 404 Ferroelectric; P = 0.26 cm−2
Y3 Al5 O12 YAG 100 Ia3d 1200
Gd3 Ga5 O12 GGG 100 Ia3d 1238
TiO2 001 P42 /mnm 459; 296 4
100
ZrO2 :Y YSZ 100 Fm3m 512 Piezoelectric
ZnO 001 P63 mc 325;521 −24
SiO2 100 P33 21 491;541 −18
Si 100 Fd3m 542 −4 Large wafers available; buffered
111 with CeO2
GaAs 001 F43m 565
905
906 J. M. D. Coey

Fig. 29 Methods for preparing thin films by a vapor or plasma condensing on a substrate:
(a) thermal evaporation, (b) e-beam evaporation, (c) pulsed-laser deposition, and (d) sputtering

from the source arrive at the substrate without collision, but at higher pressure
they are thermalized by collision with the gas atoms in the chamber. The room-
temperature mean free path of the atom λ is approximately 6/P, where λ is measured
in millimeters and P is the chamber pressure in pascals. In situ measurement of the
thicknesses of the thin films may be achieved by optical reflectometry or with a
quartz crystal monitor.

Thermal Evaporation
Resistive evaporation is a commonly used vacuum deposition process in which
electrical energy is used to heat a boat containing the charge or a filament which
heats the material to be deposited to the point of evaporation [55]. The vapor
condenses in the form of thin film on the cold substrate surface, at a rate that
depends on the inverse square of the source-target distance. The method is restricted
to materials with moderately low melting points to avoid contamination by the
boat, which is usually made of graphite, molybdenum, or tungsten. Electron beam
(e-beam) evaporation is a variant that is used both in research and on an industrial
scale for thin film coatings. Here a target of the material to be evaporated is
bombarded with a beam of high-energy electrons. The target is placed directly in
a water-cooled crucible or else in a crucible liner made from a material such as
graphite or tungsten with a high melting point that does not alloy with it. The melting
is localized, and the evaporated target material is transported to a substrate. E-beam
evaporation can also be used in a reactive environment to make oxide films by
evaporating a metal target in the presence of oxygen that is either released directly
into the evaporation chamber or else introduced via a low-energy divergent-beam
plasma ion source directed at the substrate. This technique, known as ion-beam-
assisted deposition (IBAD), produces high-quality stoichiometric films. Very high
17 Magnetic Oxides and Other Compounds 907

deposition rates can be achieved by e-beam evaporation compared to MBE or


sputtering, which is advantageous for thick films.

Molecular Beam Epitaxy (MBE)


MBE is a more sophisticated type of vacuum evaporation [17] used to lay down
layers of materials a few atoms thick in ultrahigh vacuum (UHV). Beams of the
constituent elements are generated from heated effusion cells and travel without
scattering to a substrate where they combine to form an epitaxial film. The growth
rate depends on the flux of material in the atomic or molecular beams, which can
be controlled by the evaporation rate and, most importantly, switched on and off
with shutters in a fraction of the time required to grow one monolayer. Typical
growth rates are very slow, of order a monolayer per second or a micron per hour.
MBE produces high-quality layers with very abrupt interfaces and good control of
thickness, doping, and composition, thanks to the high degree of control possible. It
is a valuable tool for developing sophisticated electronic and optoelectronic devices,
but it is not used in production.

Sputtering
Sputtering, the preferred thin film vacuum deposition technique in industry, is
widely used in research laboratories as well. Sputtered films exhibit excellent,
reproducibility, uniformity, density, purity, and adhesion. It is possible to make
oxides, nitrites, and other compounds of precise composition by reactive sputtering
from metal targets [98]. In dc sputtering, the substrates are placed in a high vacuum
chamber that is pumped down before a low pressure (0.05–1 Pa) of the process gas,
usually argon, is introduced. Sputtering starts when a negative potential of a few
hundred volts is applied to a target of the material to be deposited, causing a plasma
or glow discharge. Positively charged Ar+ ions generated in the plasma collide with
the negatively biased target. The momentum transfer ejects atomic-scale particles
from the target, which traverse the chamber and are deposited as a thin film on the
surface of the substrate. A magnetic field is usually created near the target surface by
an array of permanent magnets, known as a magnetron, which serve to improve the
ionization efficiency. Oxygen can be mixed with the argon sputtering gas to prepare
oxides from metal targets. Alternatively, oxide or other insulating films can be made
directly by radiofrequency sputtering using an rf power supply that usually operates
at 13.56 MHz. For part of the cycle, Ar ions bombard the target; while for the rest of
the cycle, electrons neutralize the buildup of positive charge. Electrons also ionize
the argon to create the plasma. An argon pressure of 0.02 Pa is usually sufficient to
maintain the radiofrequency discharge.
Most sputtering systems have multiple targets and power supplies, which permits
fabrication of the complex thin film stacks used for spin electronic devices. Double
facing targets used for off-axis deposition are particularly suitable for making
uniform, high-quality oxide thin films.
908 J. M. D. Coey

Pulsed Laser Deposition (PLD)


PLD is perhaps the most versatile physical method to deposit high-quality metallic
and insulating films in the laboratory; it has been intensively developed since 1987
[18]. The technique employs high-power UV laser pulses (typically ∼108 Wcm−2 )
from a 248 nm KrF excimer laser or a frequency tripled Nd-YAG laser to ablate
ionized material from the surface of a ceramic target in vacuum. This ablation event
produces a transient, highly luminous directional plasma plume that expands rapidly
away from the target surface. The plume deposits material onto the substrate and a
thin film is formed. The plume is stoichiometrically similar to the target, and so
thin films of roughly the same composition of the target can be easily produced.
Rastering the laser spot across the target surface or moving the substrate during
deposition yields films of uniform thickness and composition. The targets are easily
fabricated dense ceramic discs 10–25 mm in diameter. The energy of the beam
can range from 100 − 500 mJ per pulse, and typical fluences of 1–5 J cm−2 are
incident upon the target. With a repetition rate between 1 and 20 Hz. The vacuum
chamber can be partially filled with oxygen or nitrogen to allow deposition at higher
pressures.
A problem with the method was that liquid droplets or particulates may
contaminate the plume and settle on the surface of the growing film. This can be
avoided by velocity filters or off-axis deposition, where the substrate is parallel to
the plume. The method has been developed to allow for oxide films to be deposited
on large-area buffered Si substrates.

Chemical Vapor Deposition (CVD)


CVD is an industrially significant process with good stability and reproducibility,
which is used in a wide array of applications. CVD involves depositing a solid
film from a gaseous molecular precursor [99]. Different energy sources, precursor
gases, and substrates are used, depending on the desired product, but the precursors
must be volatile yet stable enough to be able to be delivered to the reactor where
the volatilized precursor (such as silane, an organometallic or a metal coordination
complex) is passed over a heated substrate. Thermal decomposition of the precursor
produces a thin film deposit, and ideally, the ligands associated with the precursor
are cleanly lost to the gas phase as reaction products. Pressure and temperature
are the important variables. A variant involving reversible conversion of nonvolatile
species into volatile derivatives is chemical vapor transport (CVT) [90] where the
volatile derivative migrates throughout a sealed reactor, typically a sealed, evacuated
quartz tube heated in a tube furnace. Because the tube is in a temperature gradient,
the volatile derivative reverts to the parent solid, which is deposited as a thin film
on a substrate, and the transport agent is released at the end opposite to where it
originated.

Atomic Layer Deposition (ALD)


ALD is based on sequential, self-limiting surface chemical reactions [46, 99].
This unique growth technique can provide atomic layer control and allow ultrathin
17 Magnetic Oxides and Other Compounds 909

conformal films to be deposited on very high aspect ratio structures. ALD deposits
films using pulses of gas that produce one atomic layer at a time. Within fairly
wide process windows, the deposited film thickness only depends on the number of
deposition cycles, providing extremely high uniformity and thickness control. ALD
reactions are typically carried out in the 200–400◦ C range. In particular, ALD is
used in the semiconductor industry for high-k gate dielectrics (HfO2 , ZrO2 ), but
with appropriate organometallic precursors, it has potential for ultrathin magnetic
oxide films.
Wet chemical methods to produce oxide thin films include dip coating and spray
pyrolysis.

Magnetic Oxide Monolayers

An important question is, how do the properties of a magnetic thin film differ from
those of the bulk material? There are obvious differences. Usually, the surfaces
of bulk material are neglected, and it is regarded as an infinite three-dimensional
continuum. The thin film is different. It is quasi two-dimensional with two quite
different surfaces – a lower one of effective thickness al in contact with the substrate
or buffer and an upper one of thickness au in contact with vacuum, the cap layer or
ambient air; al and au may be of order a nanometer, so if the total film thickness
is t, the bulk-like fraction, (t − al − au )/t, is a number that may be substantially
less than 1 or even zero for very thin films. The problem is aggravated by our
limited understanding of the real structure of the upper or lower oxide film surfaces,
especially the upper one in ambient conditions, where the classical physical methods
of surface characterization in UHV are not relevant. A monolayer of an adsorbed gas
forms in a matter of minutes, even in a vacuum of 10−5 Pa.
Surfaces of the principal iron oxides magnetite, hematite, maghemite, and
wüstite have been intensively investigated by surface scientists, and there is a
comprehensive review of their work by Parkinson [83]. They are of interest not only
for magnetism but also for catalysis, photoelectrochemistry, and nanomedicine. The
clean surfaces tend to reflect the defect chemistry of the bulk, with a close-packed
oxygen lattice maintained up to the edge and subsurface cation vacancies. Best
studied is the (100) surface of magnetite, which can readily be prepared in UHV
with a B-site iron/oxygen termination. The reconstructed surface is accompanied
by two subsurface octahedral cation vacancies and a tetrahedral interstitial, giving
a composition Fe11 O16 in the first unit cell. While bulk magnetite is half-metallic,
with only minority-spin t2g electrons at the Fermi level, there is little evidence for
complete spin polarization at the surfaces of magnetite based thin film structures.
The problems are less acute for metal thin films, but even there effects of
reduced bandwidth for the surface atoms with fewer nearest neighbours than the
bulk, reduced symmetry which can lead to strong perpendicular surface magnetic
anisotropy that scales with surface area and magnetic excitations with a two-
dimensional character may all strongly modify the magnetic ordering temperature,
and even alter the nature of the magnetic order. The nonmagnetic metals vanadium,
910 J. M. D. Coey

ruthenium, and rhodium have been reported to become ferromagnetic in layers 1 - 2


monolayers thick.
In oxides, as in metals, there are modifications of the intrinsic magnetic
properties of the material – Curie or Néel temperature, magnetization, anisotropy,
magnetostriction, and magnetoresistance. These quantities are modified by the
biaxial strain imposed by the substrate, which is gradually relaxed for thicker films.
The observation of high-field colossal magnetoresistance, much greater than that
observed previously in ceramics od single crystals, was a catalyst to investigation of
thin film magnetic oxides [109].
An additional consideration in oxides is the polar nature of the surface or
interface [49]. Assuming formal charges Sr2+ , Ti4+ , and O2− in SrTiO3 , the
sequence of planes along a <001> direction (Fig. 14a) is SrO/TiO2 / . . . . Each is
uncharged, so the (001) planes are nonpolar. However, in a <111> direction, the
stacking sequence is SrO3 /Ti/ . . . .. Each has a charge ±4e, so the (111) surfaces are
polar, and there is a net electric dipole moment in each bilayer. These add, and the
electrostatic potential and the energy in the electric field increase with each bilayer,
unless there is a compensating electric field from surface charge, needed to avoid the
energy divergence, known as the polar catastrophe. The potential step for a single
bilayer is of order volts. The stability condition to avoid the catastrophe is that the
sum of the electric fields created by the polarization and the surface charge must
be zero, which requires a surface charge of half the layer charge, ±2e in the case
of (111) SrTiO3 . Real layer charges will differ from the formal ones because of
the effects of covalency, so the charges of the (001) layers differ from zero – (001)
SrTiO3 is slightly polar – and the (111) layer charges are considerably less than ±4e.
Nevertheless, there is always some sort of electrostatic surface reconstruction at
polar oxide surfaces. It may involve subsurface cation vacancies, oxygen vacancies,
or adsorbed charged species, but in any case it is likely to influence the properties
of magnetic oxides.

Dead Layers
By measuring the magnetization Ms of an oxide film as a function of film thickness
t, and plotting Ms t as a function of t, an extrapolated dead layer t* thickness can
be deduced. Some data for (001) (La0.7 Sr0.3 )MnO3 films on different substrates
are shown in Fig. 30 [88]. This procedure does not distinguish between al and au .
Furthermore, the analysis should take account of any weak ferromagnetic signal
persisting at high temperatures that arises from the surface of the substrate. It
can be separated by measurements above the Curie temperature of the manganite.
This d-zero magnetism arising at the oxide surface has been studied in different
SrTiO3 substrates [31], but the explanation is an open question [20]. The dead layer
thicknesses on LSAT, STO, and LAO are 0.8 nm, 1.5 nm, and 3.0 nm, respectively.
It remains to be established whether the dead layer is at the manganite surface or
interface (or both) and what is the relative importance of substrate polarity and
epitaxial strain.
17 Magnetic Oxides and Other Compounds 911

Fig. 30 Plot to estimate the


dead layer thickness t* for
(La0.7 Sr0.3 )MnO3 films on
(La,Sr)(Al,Ta)O3 , SrTiO3 , or
LaAlO3 substrates [88]

Dilute Magnetic Oxides


A dilute magnetic oxide is a nonmagnetic oxide of general formula NOη where η is
an integer or rational fraction, in which a small fraction of magnetic 3d cations M
has been introduced. The formula of the dilute magnetic oxide is:

(Mx N1−x ) Oη ,

with x  0.1. From the discussion of percolation in Fig. 2, and the distribution of
magnetic ordering temperatures in Fig. 10, we would not anticipate any long-range
magnetic order in such a material when x is below the percolation threshold xp ,
which is approximately 2/Zc , where Zc is the cation-cation coordination number;
for example Zc is 12 for ZnO and 10 for TiO2 (rutile). Instead, we anticipate a
paramagnetic response reflecting the distribution of isolated ions, nearest-neighbor
pairs, and a few larger clusters, as illustrated in Fig. 2b). The dimers and clusters
are expected to couple antiferromagnetically, with a net moment for odd-membered
clusters. The paramagnetic response can be modeled approximately as a sum of
Curie and Curie-Weiss terms:

χ = C1 /T + C2 / (T–θ ) (9)

where C1 and C2 are Curie constants and θ is negative with magnitude <100 K,
representing the strength of a single antiferromagnetic exchange bond. This pre-
dicted paramagnetic behavior is found in well-crystallized bulk samples such as in
Co-doped ZnO crystals where the Co2+ substitutes for Zn on tetrahedral sites. The
Co-O-Co superexchange is antiferromagnetic.
It turns out that thin films of these oxides behave differently [21]. Following
a 2001 paper by Matsumoto et al. [70] on thin films of Co-doped anatase TiO2
produced by PLD, numerous reports appeared of ferromagnetic-like behavior in thin
films and nanocrystalline samples of dilute magnetic oxides. Some representative
results are shown in Fig. 31. Remarkably, the ferromagnetic-like behaviour was
observed at room temperature, and the corresponding Curie temperature should
912 J. M. D. Coey

Fig. 31 Some representative magnetization curves of thin film of dilute magnetic oxides [29]

be >400 K. Table 32 collects some of the early results [22]. The films may be
semiconducting (ZnO), insulating (TiO2 ), or metallic ((La,Sr)TiO3 ), and the effect
is seen for a variety of 3d cations – and even with nonmagnetic dopants such as
Sc3+ in ZnO.
At first these materials were believed to be dilute magnetic semiconductors
(DMS), where the 3d ions order ferromagnetically under the influence of long-range
exchange interactions. The thin films were thought to resemble (Ga1-x Mnx )As,
where homogeneous ferromagnetic ordering of Mn2+ ions, which occupy a level
just below the top of the As 4p valence band, occurs with a value of TC as high as
170 K for doping at the solubility limit x = 0.07 [59].
The possibility of high-temperature ferromagnetism in ZnO doped with 5% of
manganese due to a mechanism similar to that in (Ga,Mn)As was suggested by
Dietl in 2000 [36]. The ferromagnetism was mediated by spin-polarized holes in the
oxygen p band. However, ZnO, the most studied matrix [80], is exceptional insofar
as all the cations are in tetrahedral coordination and the bonding has significant
covalent character with mobile 2p holes Otherwise, oxides do not usually make good
semiconductors because the carriers tend to be trapped and form small polarons
because of their Coulomb interaction with the oxygen anions. This is the reasdon
why many off-stoichiometric oxides remain insulating.
The emerging consensus is that these dilute magnetic oxides are not dilute
magnetic semiconductors [21]. They are magnetically inhomogeneous, and only
a small fraction (of order 1% in thin films and 10 ppm in nanoparticles) of the
volume could be ferromagnetically ordered. The magnetic moment scales with
the film area rather than its volume, with values of order 10–20 μB per square
nm. The practice of relating the moment to the dopant content leads occasionally
17 Magnetic Oxides and Other Compounds 913

Table 33 Some reports of ferromagnetic oxide thin films with TC above room temperature [22]
Material Eg (eV) Substitutions Moment (μB ) per 3d ion TC (K)
TiO2 3.2 V – 5% 4.2 >400
Co – 7% 0.3 >300
Co – 1–2% 1.4 >650
Fe – 2% 2.4 300
SnO2 3.5 Fe – 5% 1.8 610
Co – 5% 7.5 650
ZnO 3.3 V – 15% 0.5 >350
Mn– 2.2% 0.16 >300
Fe5%,Cu1% 0.75 550
Co – 10% 2.0 280–300
Ni – 0.9% 0.06 >300
CeO2 2.0 Co – 3% 5.8 >750
In2 O3 3.7 Fe – 5% 1.4 >600
Cr – 2% 1.5 900
ITO 3.5 Mn – 5% 0.8 >400
(La,Sr)TiO3 – Co – 1.5% 2.5 550

to unreasonable moments per 3d ion (Table 33). The magnetism [62] (when it
is not attributable to experimental artifacts or impurity phases such as embedded
nanoclusters of metallic cobalt) is closely associated with lattice defects in the
material. Furthermore, atomic-scale probes such as XMCD [102] and Mössbauer
spectroscopy [28] failed to reveal any magnetic order of the 3d dopants at room
temperature. In dilute Co-doped ZnO films, for example, no evidence could be found
for ferromagnetic interactions among the dopants [76], and even in concentrated
films with x = 0.6, the antiferromagnetic ordering temperature does not exceed
20 K [77].
What is remarkable about the magnetization curves is that they are essentially
anhysteretic and exhibit little or no temperature dependence between 4 K and 400 K.
Curves obtained at different temperatures often superpose after correcting the data
for a diamagnetic or paramagnetic slope arising from the substrate or from the
oxide film itself. The lack of temperature dependence excludes any possibility of
superparamagnetism, for which the magnetization curves should superpose when
M/Ms is plotted versus (H/T).
The absence of hysteresis is a further argument against uniform magnetization, as
would be expected for a DMS. The small net magnetization values Ms ≈ 10 kA m−1
inferred for thin films should lead to an appreciable anisotropy field, with a uniaxial
anisotropy constant due to intrinsic crystal field or substrate-induced strain, and
therefore produce coercivity at low temperature. Another serious problem is the high
value of the Curie temperature. Relatively few insulating oxides are ferromagnetic
(Fig. 10); they usually order antiferromagnetically or ferrimagnetically on account
of the antiferromagnetic superexchange. Almost none order magnetically above
1000 K. It is unreasonable to expect oxide films with just 5% of magnetic ions to
have a Curie temperature exceeding 100 K. The magnetic energy density varies as x
914 J. M. D. Coey


or x, so high Curie temperatures could not be achieved with uniformly magnetized
material.
A defect-related origin rather than a magnetic dopant-related origin of the fer-
romagnetism is supported by its observation in some samples of undoped material,
including films of HfO2 , ZnO, TiO2 , and In2 O3 . This phenomenon is known as
d0 magnetism [20], and it reinforces the idea that while defects are necessary, 3d
dopants are not essential. It is a challenge to associate the magnetism, which is often
not quantitatively reproducible, with a specific surface, interface, or grain-boundary
defects. Extended defects may be a more likely source of percolating magnetism
than point defects. As regards the origin of d-zero magnetism two quite different
ideas are in contention.
One is to model the defects, probably oxygen vacancies, in terms of a narrow,
spin-split defect-related impurity band. This can lead to a high Curie temperature
since spin-wave excitations are suppressed in the half-metallic band. It is a Stoner
model of the magnetism, involving a small fraction of the sample volume. The
transition-metal dopants may nonetheless play a role when they exist in different
charge states, providing a reservoir of charge for the impurity band to ensure filling
to the point where the Stoner criterion is satisfied. This model is known as charge-
transfer ferromagnetism [28].
An alternative view is that there is actually no spin ferromagnetism in these
oxide films. The reversible nonlinear response to the magnetic field is really an
unusual form of giant orbital paramagnetism, which has been related to the coherent
response of quasi-two-dimensional electron systems to fluctuations of the vacuum
electromagnetic field [32]. The definitive explanation remains an open question [20].

Oxide Heterostructures and Interfaces

Ultrathin nonmagnetic oxide barriers play an important role in spin electronics [7],
which has become the most innovative area of applied magnetism in the twenty-
first century [113]. Early tunnel barriers were made of amorphous aluminum oxide,
but the discovery that crystalline (100) MgO layers act as directional spin filters
when grown on a (100) film of αFe, or a body-centered cubic Fe-Co alloy [82, 115],
led to magnetic tunnel junctions (Fig. 28b) with a magnetoresistance in excess of
200%, which made them attractive both as magnetic sensors and memory or logic
elements. Other nonmagnetic oxide tunnel barriers, such as spinel MgAl2 O4 , have
been developed, but none rivals the efficiency of MgO.
Thin films of magnetic oxides and sulfides such as ferromagnetic EuO or EuS
and ferrimagnetic CoFe2 O4 or NiFe2 O4 can serve as tunnel spin filters with a
nonmagnetic metal electrode, since the bandgap is different for minority- and
majority-spin electrons [41].
Spin electronics [113] has progressed from its initial emphasis on magnetore-
sistive sensors for magnetic recording [119] to the development of a competitive
nonvolatile magnetic memory and logic technology based on switching magnetic
17 Magnetic Oxides and Other Compounds 915

tunnel junctions. At first the switching was based on current-generated magnetic


fields, but that approach proved not to be scalable. Alternative approaches were
then discovered using the angular momentum in a spin-polarized current to switch
the bistable magnetic element directly (spin transfer torque) or indirectly by spin
injection from a spin current in an adjacent heavy metal layer such as Ta, W, or
Pt, making use of the spin Hall effect (spin-orbit torque). Magnetic oxides have so
far played little role here, the best ferromagnetic layers being iron, cobalt, and their
alloys.
A long-term ambition has been to devise a form of oxide spin electronics, where
memory or logic operations can be conducted in all-oxide thin film stacks grown
on large-scale silicon wafers [8]. An attractive prospect is the use of electric fields
(or strain) rather than electric currents to control magnetic switching in stacks with
adjacent ferromagnetic (or ferrimagnetic) and ferroelectric (or piezoelectric) layers,
thereby reducing the rapidly escalating energy cost of the big data revolution.
Growth of oxide heterostructures is facilitated by the structure of the basic
oxide scaffold, which is often close-packed, or else has a perovskite-type struc-
ture. The choice of conducting ferromagnetic oxides is quite limited: CrO2 ,
(La0.7 Sr0.3 )MnO3 , Fe3 O4 , and some double perovskites order magnetically well
above room temperature, but a drawback of any oxide magnet is that its mag-
netization is less than a third that of iron. This is unavoidable because most
of the unit cell volume is occupied by the large nonmagnetic O2− anions, with
high-spin ferric iron Fe3+ or other magnetic ions in the interstices. Ferrimagnetic
structures reduce the magnetization further. Interfacial dead layers tend to degrade
the performance of the magnetic oxide layers, and the room-temperature candidate
ferro- or ferrimagnetic oxides do not perform nearly as well as their metallic coun-
terparts. However, SrRuO3 has played a part in low-temperature proof-of-principle
perovskite-structure stacks when a good spin-polarized ferromagnetic metal was
required. An advantage of all-oxide stacks is that ferroelectric layers such as BiTiO3
or BiFeO3 , a rare room-temperature multiferroic, can be readily incorporated into
the stacks. The cycloidal character of the weak ferromagnetism is suppressed in thin
films of BiFeO3 , and it then exhibits a large linear magnetoelectric effect, of order
2 nsm−1 . There is a prospect of using magnetoelectric switching combined with
spin-orbit detection of state in new ultralow energy nonvolatile logic and memory
beyond CMOS technology [69].

Oxide Interfaces
At interfaces between oxides, the three-dimensional symmetry of each oxide is
broken, and proximity effects may emerge [30]. Interfacial exchange can modify
the magnetic structure at the interface, if one or both of the oxides is magnetically
ordered. Exchange bias, for example, may be a consequence. When the electronic
structure of the two oxides differs, charge transfer may result in an interface with a
different magnetic structure or different transport properties [30, 106].
The interfaces between polar and nonpolar oxides are especially interesting.
Charged two-dimensional surfaces are unstable in nature unless there is nearby
916 J. M. D. Coey

compensating charge, because of the energy density ½ε0 E2 associated with the
uniform perpendicular electric field created in the surrounding space – an instability
known as the polar catastrophe. Inevitably some electronic or ionic reconstruction
must occur to resolve the problem.
A celebrated and much-studied example is the (100) interface between formally
nonpolar SrTiO3 and polar LaAlO3 [79]. Both are insulators, yet a conducting
two-dimensional electron gas (2DEG) forms just below the interface in the SrTiO3
provided the thickness of the LaAlO3 overlayer exceeds four unit cells. The electron
density and orbital character in the Ti dxy or dyz/zx bands can be modified by gating,
and there is evidence of inhomogeneous ferromagnetic order in the 2DEG at liquid
helium temperatures. Spin diffusion lengths can be on the order of a micron [106].

Spin Pumping
The technique of spin pumping has been developed to create spin currents perpen-
dicular to the interface between ferro- or ferrimagnetic and nonmagnetic layers. It is
related to the spin Hall effect, whereby a spin current is created perpendicular to a
thin film of a heavy metal such as Ta, W, or Pt by passing an electric current through
it. The spin current can modify the magnetic state of an adjacent ferromagnetic
layer. In spin pumping, Larmor precession in the magnetic layer is excited by
a radiofrequency magnetic field at a frequency determined by the ferromagnetic
resonance that depends on an applied field or the anisotropy field of the layer. In that
case, a rectified spin current is injected into the adjacent nonmagnetic layer, which
may be detected by the inverse spin Hall effect. For this to work well, a magnetic
material with very low damping is required, which makes yttrium-iron garnet the
preferred option. The quality of the interface and the oxygen stoichiometry of the
YIG are critical for efficient spin injection [114].

Magnonics
Magnonics is the magnetic transport, storage, and processing of information using
spin waves. See  Chap. 6, “Spin Waves”. The rich physics of the linear and
nonlinear dynamics of multimode spin wave systems has been extensively inves-
tigated using yttrium-iron garnet strip lines [93]. The uniquely low damping in YIG
allows spin wave propagation to be observed over distances of up to a centimeter,
a thousand times longer than in a metal such as permalloy and far greater than
the spin diffusion lengths encountered in spin electronics. Magnons, like diffusive
spin currents, involve the transport of angular momentum, which in the case of
magnons is correlated precession of the localized spin moments on neighboring
sites. The key functions of propagation, storage, amplification, echo, and logic
have all been demonstrated. Magnonics has the advantage of low power dissipation
and may be combined with spin transfer torque to enable controlled generation
of magnons along well-defined propagation directions [57]. Disadvantages are the
large footprint of the devices (square microns) and the low charge ↔ magnon
conversion efficiency at the input and output.
17 Magnetic Oxides and Other Compounds 917

Conclusion

Oxides continue to rival metals in practical importance as magnetic materials, and


they have played a critical role in developing our understanding of collective mag-
netic order with localized electrons. Transition-metal oxide systems are especially
useful for exploring the effects of frustration and spin-orbit coupling on magnetic
order. Physical and chemical constraints of crystal structure and orbital overlap
set limits on what is magnetically feasible and the associated parameters such as
magnetization, anisotropy, Curie temperature, and magnetostriction. Traditionally,
the space of achievable magnetism has been explored in bulk oxide solid solutions,
leading to optimized functional materials – Fig. 19 showed a good example. In the
future, strain and interface engineering in thin films may hold the key to the control
of the magnetism of thin films or multilayers and potentially generate original oxide-
based functionality for spin electronics.

Acknowledgments The author is grateful to Science Foundation Ireland for continued support,
including contracts 10/IN.1/I3006, 13/ERC/I2561 and 16/IA/4534.

References
1. Anderson, P.W.: Exchange in insulators: Superexchange, direct exchange and double
exchange: Ch2. In: Rado, G.T., Suhl, H. (eds.) Magnetism I, pp. 25–83. Academic Press,
New York (1963)
2. Artman, J.O., Murphy, J.C., Foner, C.S.: Magnetic anisotropy in antiferromagnetic corundum-
type sesquioxides. Phys. Rev. 138, A912 (1965)
3. Ballet, O., Coey, J.M.D., Mangin, P., Townsend, M.G.: Ferrous talc - a planar antiferromagnet.
Solid State Comm. 55, 787–790 (1985)
4. Ballet, O., Fuess, H., Wacker, K., Untersteller, E., Treutmann, W., et al.: Magnetic measure-
ments of the synthetic olivine single crystals A2 SiO4 with a = Fe, co or Ni. J. Phys. Condens.
Matter. 1, 4955–4979 (1989)
5. Ballhausen, C.J.: Introduction to Ligand Field Theory. McGraw Hill, New York (1962)
6. Bender, G., Koch, C., Oxborrow, C.A., Mørup, S., Madsen, M.B., Quinn, A.J., et al.: Magnetic
properties of feroxhyte (δ-FeOOH). Phys. Chem. Miner. 22, 333–341 (1995)
7. Bibes, M., Villegas, J.E., Barthélémy, A.: Ultrathin oxide films and interfaces for electronics
and spintronics. Adv. Phys. 60, 5 (2011)
8. Blank, D.H.A., Dekkers, M., Rijnders, G.: Pulsed laser deposition in Twente: from research
tool towards industrial deposition. J. Phys. D. Appl. Phys. 47, 34006 (2014)
9. Bonnenberg, D., Boyd, E.L., Calhoun, B.A., Folen, V.J., Gräper, W., et al.: Landholdt-
Börnstein Numerical Data and Functional Relationships in Science and Technology New
Series Group III, Vol 4 a and b, Magnetic and Other Properties of Oxides and Related
Compounds. Springer, Berlin (1970)
10. Brabers, V.A.M.: Progress in spinel ferrite research, Ch 3. In: Buschow, K.H.J. (ed.)
Handbook of Ferromagnetic Materials, vol. 8, pp. 189–324. Elsevier, Amsterdam (1995)
11. Bramwell, S.T., Gingras, M.J.P.: Spin ice state in frustrated magnetic pyrochlore materials.
Science. 294, 1495–1501 (2001)
12. Brodsky, M.B.: Magnetic properties of the actinide elements and their metallic compounds.
Rep. Prog. Phys. 41, 1548 (1978)
918 J. M. D. Coey

13. Burns, R.G.: Mineralogical Applications of Crystal Field Theory, 2nd edn. Cambridge
University Press, London (1993)
14. Cao, G., McCall, S., Shepard, M., Crow, J.E., Guertin, R.P.: Thermal, magnetic and transport
properties of single-crystal Sr1−x Cax RuO3 (0 < x < 1.0). Phys. Rev. B. 56, 321–329 (1997)
15. Castelnovo, C., Moessner, R., Sondhi, S.L.: Magnetic monopoles in spin ice. Nature. 451,
42–45 (2009)
16. Catalano, S., Gibert, M., Fowlie, J., ´iñiguez, J., Triscone, J.-M., et al.: Rare-earth nickelates
RNiO3 : thin films and heterostructures. Rep. Prog. Phys. 81, 046501 (2018)
17. Cho, A.Y., Arthur, J.R.: Molecular beam epitaxy. Prog. Solid State Chem. 10, 157 (1975)
18. Christen, H.M., Eres, G.: J. Phys. Conden. Mat. 20, 264005 (2008)
19. Coey, J.M.D.: Hard magnetic materials – a perspective. IEEE Trans. Magnetics. 47, 4671–
4681 (2011)
20. Coey, J.M.D.: Magnetism in d0 oxides. Nat. Mater. 18 (2019)
21. Coey, J.M.D.: Magnetism of dilute oxides, Ch 10. In: Tsymbal, E.Y., Zutic, I. (eds.) Handbook
of Spin Transport and Magnetism, 2nd edn. CRC Press, Boca Raton (2019)
22. Coey, J.M.D., Chambers, S.A.: Oxide dilute magnetic semiconductors – fact or fiction? MRS
Bull. 33, 1053 (2008)
23. Coey, J.M.D., Ghose, S.: Magnetic phase transitions in silicate minerals, Ch 9. In: Ghose,
S., Coey, J.M.D., Salje, E. (eds.) Structural and Magnetic Phase Transitions in Minerals, pp.
162–184. Springer, New York (1988)
24. Coey, J.M.D., Readman, P.W.: New spin structure in an amorphous ferric gel. Nature. 246,
476–478 (1973)
25. Coey, J.M.D., Smith, P.A.I.: Magnetic nitrides. J. Magn. Magn. Mater. 200, 405–424 (1999)
26. Coey, J.M.D., Barry, A., Brotto, J.-M., Rakoto, H., Brennan, S., et al.: Spin flop in goethite.
J. Phys. 7, 759–768 (1995)
27. Coey, J.M.D., Viret, M., von Molnar, S.: Mixed-valence manganites. Adv. Phys. 48, 167–293
(1999)
28. Coey, J.M.D., Stamenov, P., Gunning, R.D., Venkatesan, M., Paul, K.: Ferromagnetism in
defect-ridden oxides and related materials. New J. Phys. 12, 053025 (2010)
29. Coey, J.M.D., Venkatesan, M., Xu, H.-J.: Introduction to magnetic oxides Ch 1. In: Ogale,
S.B., Venkatesan, T.V., Blamire, M.G. (eds.) Functional Magnetic Oxides, pp. 1–30. Wiley-
VCH, Weinheim (2013)
30. Coey, J.M.D., Ariando, Pickett, W.E.: Magnetism at the edge; new phenomena at oxide
interfaces. Mater. Res. Bull. 38, 1040–1047 (2013)
31. Coey, J.M.D., Venkatesan, M., Stamenov, P.: Surface magnetism of strontium titanate. J. Phys.
Condens. Matter. 28, 485001 (2016)
32. Coey, J.M.D., Ackland, K., Venkatesan, M., Sen, S.: Collective magnetic response of CeO2
nanoparticles. Nat. Phys. 12, 694–699 (2016)
33. Connolly, T.F., Copenhaver, E.D.: Bibliography of Magnetic Materials and Magnetic Transi-
tion Temperatures. Springer, New York (1972)
34. Corliss, L., Elliott, N., Hastings, J.: Magnetic structures of the polymorphic forms of
manganous sulfide. Phys. Rev. 104, 924–928 (1956)
35. Cornell, R.M., Schwertmann, U.: The Iron Oxides; Structure, Properties, Reactions, Occur-
rence and Uses. VCH, Weinheim (2003)
36. Dietl, T., Ohno, H., Matsukura, F., Cibert, J., Ferrand, D.: Zener model description of
ferromagnetism in zinc-blende magnetic semiconductors. Science. 287, 1019 (2000)
37. Dijkkamp, D., Venkatesan, T., Ogale, X.-D.W.S.B., Inam, A., et al.: Preparation of Y-Ba-Cu
oxide superconductor thin films using pulsed laser evaporation from high Tc bulk material.
Appl. Phys. Lett. 51, 619 (1987)
38. Dionne, G.F.: Magnetic Oxides. Springer, New York (2009)
39. Dunlop, D.J., Özdemir, Ö.: Rock Magnetism: Fundamentals and Frontiers. Cambridge
University Press, Cambridge (1997)
40. Dzyaloshinsky, I.: A thermodynamic theory of “weak” ferromagnetism of antiferromagnetics.
J. Phys. Chem. Solids. 4, 241–255 (1958)
17 Magnetic Oxides and Other Compounds 919

41. Esaki, L., Stiles, P.J., von Molnar, S.: Magnetointernal field emission in junctions of magnetic
insulators. Phys. Rev Lett. 19, 852 (1967)
42. Eschenfelder, A.H.: Magnetic Bubble Technology. Springer, Berlin (1980)
43. Frederichs, T., von Dobeneck, T., Bleil, U., Dekkers, M.J.: Towards the identification
of siderite, rhodochrosite and vivianite in sediments by their low-temperature magnetic
properties. Phys. Chem. Earth. 28, 669–679 (2003)
44. Fuess, H., Ballet, O., Lottermoser, W.: Magnetic phase transitions in olivines, Ch 10. In:
Ghose, S., Coey, J.M.D., Salje, E. (eds.) Structural and Magnetic Phase Transitions in
Minerals, pp. 185–201. Springer, New York (1988)
45. Gardner, J.S., Gingras, M.J.P., Greedan, J.E.: Magnetic pyrochlore oxides. Rev. Mod. Phys.
82, 53–107 (2010)
46. George, S.M.: Atomic layer deposition: an overview. Polymer. 1550, 125 (2010)
47. Gilleo, M.A.: Ferromagnetic insulators: garnets, Ch 1. In: Wohlfarth, E.P. (ed.) Handbook of
Ferromagnetic Materials, vol. 2, pp. 1–54. Amsterdam, North Holland (1980)
48. Gong, C., Zhang, X.: Two dimensional magnetic crystals and emergent heterostructure
devices. Science. 363 (2019)
49. Goniakowski, J., Finocchi, F., Noguera, C.: Polarity of oxide surfaces and nanostructures.
Rep. Prog. Phys. 71, 016501 (2008)
50. Goodenough, J.B.: Magnetism and the Chemical Bond. Wiley-Interscience, New York (1963)
51. Greaves, C.: A powder neutron diffraction investigation of vacancy ordering and covalence in
γFe2 O3 . J. Solid State Chem. 49, 325 (1983)
52. Greenwood, N.N., Gibb, T.C.: Mössbauer Spectroscopy. Chapman and Hall, London (1971)
53. Hazen, R.M., Jeanloz, R.: Wüstite (Fe1-x O): a review of its defect structure and physical
properties. Rev. Geosci. Space Sci. 22, 37 (1984)
54. Hemberger, J., Krimmel, A., Kurz, T., Krug von Nidda, H.-A., Ivanov, V.Y., et al.: Structural,
magnetic, and electrical properties of single-crystalline La1−x Srx MnO3 (0.4 < x < 0.85). Phys.
Rev. B. 66, 094410 (2002)
55. Hill, R.J.: Physical vapour deposition, 2nd edn. Temescal, Berkley (1986)
56. Hill, A.H., Jiao, F., Bruce, P.G., Harrison, A., Kockelmann, W., Ritter, C.: Neutron diffraction
study of mesoporous and bulk hematite, r-Fe2O3. Chem. Mater. 20, 4891–4899 (2008)
57. Hoffmann, A., Bader, S.D.: Opportunities at the frontiers of spintronics. Phys. Rev. Appl. 4,
047001 (2015)
58. Hurlbut, C.S., Klein, C.: Manual of Mineralogy, 19th edn. Wiley, New York (1977)
59. Jungwirth, T., Wang, K.Y., Mašek, J., Edmonds, K.W., König, J., et al.: Prospects for high
temperature ferromagnetism in (Ga, Mn) as semiconductors. Phys. Rev. B. 72, 165204 (2005)
60. Kalid, M., Setzer, A., Ziese, M., Esquinazi, P., Spemann, D., et al.: Ubiquity of ferromagnetic
signals in common diamagnetic oxides. Phys. Rev. B. 81, 214414 (2010)
61. Kanamori, J.: Anisotropy and magnetostriction of ferromagnetic and antiferromagnetic
materials, Ch 4. In: Rado, G.T., Suhl, H. (eds.) Magnetism I, pp. 127–203. Academic Press,
New York (1963)
62. Khomski, D.I.: Transition Metal Compounds. Cambridge University Press, Cambridge (2014)
63. Kojima, H.: Fundamental properties of hexagonal ferrites with the magnetoplumbite structure,
Ch. 5. In: Wohlfarth, E.P. (ed.) Handbook of Ferromagnetic Materials, vol. 3, pp. 305–391.
Amsterdam, North Holland (1982)
64. Krupicka, S., Novak, P.: Oxide spinels, Ch 4. In: Wohlfarth, E.P. (ed.) Handbook of
Ferromagnetic Materials, vol. 3, pp. 189–303. North Holland, Amsterdam (1982)
65. Kuneš, J., Anisimov, V.I., Skornyakov, S.L., Lukoyanov, A.V., Vollhardt, D.: NiO: correlated
band structure of a charge-transfer insulator. Phys. Rev. Lett. 99, 156404 (2007)
66. Lester, C., Ramos, S., Perry, R.S., Croft, T.P., Bewley, R.I., et al.: Field-tunable spin-density-
wave phases in Sr3 Ru2 O7 . Nat. Mater. 14, 373–378 (2015)
67. Lu, H.-C., Chamorro, J.R., Wan, C., McQueen, T.M.: Universal single-ion physics in spin-
orbit-coupled d5 and d4 ions. Inorg. Chem. 57, 14443–14449 (2018)
68. Mackenzie, A.P., Maeno, Y.: The superconductivity of Sr2 RuO4 and the physics of spin-triplet
pairing. Rev. Mod. Phys. 75, 657–712 (2003)
920 J. M. D. Coey

69. Manipatruni, S., Nikonov, D.E., Lin, C.-C., Gosavi, T.A., Liu, H.-C., et al.: Scalable energy-
efficient magnetoelectric spin-orbit logic. Nature. 565, 35–43 (2019)
70. Matsumoto, Y., Murakami, M., Shono, T., Hasegawa, T., Fukumura, T., et al.: Room-
temperature ferromagnetism in transparent transition metal-doped titanium dioxide. Science.
291, 854 (2001)
71. McGuire, M.A.: Crystal and magnetic structures in layered transition metal dihalides.
Crystals. 7, 120–145 (2017)
72. Mikhaylovskiy, R.V., Hendry, E., Secchi, A., Mentink, J.H., Eckstein, M., et al.: Ultrafast
optical modification of exchange interactions in iron oxides. Nat. Commun. 6, 8190 (2015)
73. Moriya, T.: Anisotropic superexchange interaction and weak ferromagnetism. Phys. Rev. 120,
91–98 (1960)
74. Morrish, A.H.: Canted Antiferromagnetism: Hematite. World Scientific, Singapore (1994)
75. Mott, N.F.: Metal-Insulator Transitions, 2nd edn. Taylor and Francis, London (1990)
76. Ney, A., Ollefs, K., Ye, S., Kammermeier, T., Ney, V., et al.: Absence of intrinsic ferromag-
netic interactions of isolated and paired co dopant atoms in Zn1-x Cox O with high structural
perfection. Phys. Rev. Lett. 100, 157201 (2008)
77. Ney, V., Henne, B., Lumetzberger, J., Wilhelm, F., Ollefs, K., et al.: Coalescence-driven
magnetic order of the uncompensated antiferromagnetic Co doped ZnO. Phys. Rev. B. 94,
224405 (2016)
78. Obradors, X., Solans, X., Collomb, A., Samaras, D., et al.: Crystal structure of strontium
hexaferrite. J. Solid State Chem. 72, 218–224 (1988)
79. Ohtomo, A., Hwang, H.Y.: A high-mobility electron gas at the LaAlO3 /SrTiO3 heterointer-
face. Nature. 427, 423–426 (2004)
80. Pan, F., Song, C., Liu, X.-J., Yang, Y.-C., Zeng, F.: Ferromagnetism and possible application
in spintronics of transition-metal-doped ZnO films. Mater. Sci. Eng. Reports. 62, 1 (2008)
81. Park, J.H., Vescovo, E., Kim, H.J., Kwon, C., Ramesh, R., et al.: Direct evidence for a half-
metallic ferromagnet. Nature. 392, 794 (1998)
82. Parkin, S.S.P., Kaiser, C., Panchula, A., Rice, P.M., Hughes, B., et al.: Giant tunneling
magnetoresistance with MgO (100) tunnel barriers. Nat. Mater. 3, 862–867 (2004)
83. Parkinson, G.S.: Iron oxide surfaces. Surf. Sci. Reports. 71, 272–365 (2016)
84. Pearce, C.I., Pattrick, R.A.D., Vaughan, D.J.: Electric and magnetic properties of sulphides.
Rev. Mineral. Geochem. 61, 127–180 (2006)
85. Pegg, J.T., Aparicio-Angles, X., Storr, M., de Leeuw, N.H.: DFT + U study of the structures
and properties of the actinide dioxides. J. Nucle. Mater. 492, 269–278 (2017)
86. Penney, T., Shafer, M.W., Torrance, J.B.: Insulator-metal transition and long-range magnetic
order in EuO. Phys. Rev. B. 5, 3669 (1972)
87. Philipp, J.B., Majewski, P., Alff, L., Erb, A., Gross, R., et al.: Structural and doping effects
in half-metallic double perovskites A2 CrWO6 (A=Sr, Ba, and Ca) Phys. Rev. B. 68, 144431
(2003)
88. Porter, S.B., Venkatesan, M., Dunne, P., Doudin, B., Rode, K., et al.: Magnetic dead layers in
La0.7 Sr0.3 MnO3 revisited. IEEE Trans. Magn. 56, 6000904 (2017)
89. Pullar, R.C.: Hexagonal ferrites a review of the synthesis, properties and applications of
hexaferrite ceramics. Prog. Mater. Sci. 57, 1191–1334 (2012)
90. Schaefer, H.: Chemical Transport Reactions. Academic Press, New York (1964)
91. Schwertmann, U., Murad, E.: Effects of pH on the formation of goethite and hematite from
ferrihydrite. Clays Clay Miner. 31, 277 (1983)
92. Sen, M.S., Wright, J.P., Attfield, J.P.: Charge order and three-site distortions in the structure
of magnetite. Nature. 481, 173 (2011)
93. Serga, A.A., Chumak, A.V., Hildebrands, B.: YIG magnonics. J. Phys. D. Appl. Phys. 43,
264002 (2010)
94. Skomski, R., Coey, J.M.D.: Permanent Magnetism, pp. 257–264. IOP Publishing, Bristol
(1999)
17 Magnetic Oxides and Other Compounds 921

95. Skomski, R., Coey, J.M.D.: Magnetic anisotropy – how much is enough for a permanent
magnet? Scripta Mater. 112, 3–8 (2016)
96. Smit, J., Wijn, H.P.J.: Ferrites. Philips Technical Library, Eindhoven (1959)
97. Stäblein, H.: Hard ferrites and plastoferrites, Ch. 7. In: Wohlfarth, E.P. (ed.) Handbook of
Ferromagnetic Materials, vol. 3, pp. 441–601. Amsterdam, North Holland (1982)
98. Steinbeiss, E.: Ch 13. In: Ziese, M., Thornton, M.J. (eds.) Spin Electronics, pp. 296–315.
Springer Verlag, Berlin (2001)
99. Sze, S.M., Lee, M.K.: Semiconductor Devices: Physics and Technology, Ch 12, 3rd edn.
Wiley, New York (2012)
100. Takayama, T., Kato, A., Dinnebier, R., Nuss, J., Kono, H., et al.: Hyperhoneycomb iridate
β-Li2 IrO3 as a platform for Kitaev magnetism. Phys. Rev. Lett. 114, 077202 (2015)
101. Tanabe, Y., Sugano, S.: On the absorption spectra of complex ions I – III. J. Phys. Soc. Jpn.
9, 753 (1954).; 9 766 (1954); 11 864 (1956)
102. Tietze, T., Gacic, M., Schütz, G., Jacob, G., Brück, S., et al.: XMCD studies on Co and Li
doped ZnO magnetic semiconductors. New J. Phys. 10, 055009 (2008)
103. Treves, D.: Studies on orthoferrites at the Weizmann Institute of Science. J. Appl. Phys. 36,
1033–1039 (1965)
104. Valenzuela, R.: Magnetic Ceramics. Cambridge University Press, Cambridge (1994)
105. van Stapele, R.P.: Sulphospinels, Ch 8. In: Wohlfarth, E.P. (ed.) Ferromagnetic Materials, vol.
3, pp. 607–745. Amsterdam, North Holland (1982)
106. Varignon, J., Vila, L., Barthélemy, A., Bibes, M.: A new spin for oxide interfaces. Nat. Phys.
14, 322–325 (2018)
107. Vasala, S., Karppinen, M.: A2 B BO6 perovskites: a review. Prog. Solid State Chem. 43, 1–36
(2015)
108. Verwey, E.J.W.: Electronic conduction of magnetite (Fe3 O4 ) and its transition point at low
temperature. Nature. 144, 327 (1937).; Verwey E.J.W., Haayman, P.W. Electronic conduction
and transition point of magnetite (Fe3 O4 ). Physica 8, 979 (1941)
109. von Helmolt, R., Wecker, J., Holzapfel, B., Schultz, L., Samwer, K.: Giant negative
magnetoresistance in perovskite-like La2/3 Ba1/3 MnOx ferromagnetic films. Phys. Rev. Lett.
71, 2331 (1993)
110. Walz, F.: The Verwey transition, − a topical review. J. Phys. Cond. Matter. 14, R285 (2002)
111. Wang, Z.-S., Oureshi, N., Yasin, S., Mukhin, A., Ressouche, E., et al.: Magnetoelectric effect
and phase transitions in CuO in external magnetic fields. Nat. Comm. 7, 10295 (2016)
112. Wasey, A.H.M.A., Karmakar, D., Das, G.P.: Manifestation of long-range ordered state in VX2
(X = Cl, Br, I) systems. J. Phys. Condens. Matter. 25, 476001 (2013)
113. Xu, Y.-B., Awschalom, D.D., Nitta, J. (eds.): Handbook of Spintronics. Springer, Berlin
(2015)
114. Yang, F.-Y., P. C.: Hammel: FMR-driven spin pumping in Y3 Fe5 O12 -based structures. J. Phys.
D. Appl. Phys. 51, 253001 (2018)
115. Yuasa, S., Nagahama, T., Fukushima, A., Suzuki, Y., Ando, K.: Giant room-temperature
magnetoresistance in single-crystal Fe/MgO/Fe magnetic tunnel junctions. Nat. Mater. 3,
868–871 (2004)
116. Zener, C.: Interactions between the d-shells in the transition metals. Phys. Rev. 81, 40; 82 403
(1951)
117. Zhang, Z., Satpathy, S.: Electron states, magnetism and the Verwey transition in magnetite.
Phys. Rev. B. 44, 13319 (1991)
118. Zhao, D.-P., Zhang, L.-G., Malik, I.A., Liao, M.-H., Cui, W.-Q., et al.: Finite temperature
magnetism of CrSe and CrTe. Nano Res. 11, 3116–3121 (2018)
119. Ziese, M., Thornton, M.J.: Spin Electronics Lecture Notes on Physics, vol. 569. Springer,
Berlin (2001)
922 J. M. D. Coey

Michael Coey received his PhD from the University of Manitoba


in 1971. He has worked at the CNRS, Grenoble, IBM, Yorktown
Heights, and, since 1979, at Trinity College Dublin. Author
of several books and many papers, his interests include amor-
phous and disordered magnetic materials, permanent magnetism,
oxides and minerals, d0 magnetism, spin electronics, magneto-
electrochemistry, magnetofluidics, and the history of ideas.
Dilute Magnetic Materials
18
Alberta Bonanni , Tomasz Dietl , and Hideo Ohno

Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 926
Dilute Magnetic Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 928
Dilute Magnetic Metals (DMMs) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 928
p-Type Dilute Ferromagnetic Semiconductors (DFSs) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 928
Other Dilute Ferromagnetic Semiconductors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 929
Dilute Magnetic Semiconductors (DMSs) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 929
Dilute Magnetic Topological Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 930
Heterogenous Magnetic Semiconductors and Oxides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 930
Energy States of Magnetic Dopants in Solids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 931
Exchange Interactions Between Band and Localized Spins . . . . . . . . . . . . . . . . . . . . . . . . . . . 936

A. Bonanni
Institut für Halbleiter- und Festkörperphysik, Johannes Kepler University, Linz, Austria
e-mail: alberta.bonanni@jku.at
T. Dietl
International Research Centre MagTop, Institute of Physics, Polish Academy of Sciences,
Warsaw, Poland
WPI Advanced Institute for Materials Research, Tohoku University, Sendai, Japan
e-mail: dietl@ifpan.edu.pl
H. Ohno ()
WPI Advanced Institute for Materials Research, Tohoku University, Sendai, Japan
Laboratory for Nanoelectronics and Spintronics, Research Institute of Electrical Communication,
Tohoku University, Sendai, Japan
Center for Spintronics Integrated System, Tohoku University, Sendai, Japan
Center for Innovative Integrated Electronic Systems, Tohoku University, Sendai, Japan
Center for Science and Innovation in Spintronics (Core Research Cluster), Tohoku University,
Sendai, Japan
Center for Spintronics Research Network, Tohoku University, Sendai, Japan
e-mail: ohno@riec.tohoku.ac.jp

© Springer Nature Switzerland AG 2021 923


J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_21
924 A. Bonanni et al.

Effects of sp-d(f ) Exchange Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 938


Spin-Splitting of Extended States: Weak Coupling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 938
Spin-Splitting of Extended States: Strong Coupling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 940
Alloy and Spin-Disorder Scattering: Weak Coupling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 941
Alloy and Spin-Disorder Scattering: Strong Coupling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 942
Bound Magnetic Polarons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 942
Quantum Localization and Mesoscopic Phenomena: Colossal
Magnetoresistance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 945
Interplay of sp–d(f ) Exchange Interactions and Spin-Orbit Coupling . . . . . . . . . . . . . . . . 945
Dominant Spin-Spin Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 946
Dipole-Dipole Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 946
Direct Spin-Spin Interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 946
Superexchange . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 947
Carrier-Mediated Spin-Spin Coupling: Intra- and Interband Contributions . . . . . . . . . . . . 949
Magnetic Properties of Dilute Magnetic Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 952
Spin-Glass Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 952
p-Type Dilute Ferromagnetic Semiconductors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 957
Dilute Ferromagnetic Insulators and Topological Insulators . . . . . . . . . . . . . . . . . . . . . . . . 960
Heterogenous Magnetic Semiconductors and Oxides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 964
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 968

Abstract

This chapter reviews properties of dilute magnetic metals, semiconductors, and


topological materials, with a focus on compound semiconductors. Comprehen-
sive studies of these systems have bridged the physics and functionalities of
semiconductors and magnetic materials as well as opened the door for novel
concepts of spintronic devices by demonstrating the control of magnetization and
spin currents by light, strain, electric fields, and currents in both topologically
trivial and nontrivial systems (chapter “SSE”). Furthermore, dilute magnetic
materials have gained a status of model systems for assessing the nature of
thermodynamic phase transformations in the presence of spatial disorder, com-
peting interactions, and frustration. In particular, studies of spin-glass freezing
have challenged standard theoretical models of statistical physics and triggered
the development of new conceptual and computational tools impacting presently
various fields of science and computer hardware. At the same time, progress
in controlling magnetic ion distribution at the nanoscale results in embedded
nanostructures that have a potential for applications in electronics, spintronics,
photovoltaics, plasmonics, and thermoelectrics.

List of Symbols

A Spectral density of states


aB Effective Bohr radius
bq Fourier transform of electron density
F Free energy density
F Free energy density functional
 Mean free path
N Charge state
18 Dilute Magnetic Materials 925

N0 Cation density
p Hole carrier concentration
P Spin polarization
P Probability distribution
Tf Spin-glass freezing temperature
u Periodic part of the Bloch function
Vkd Hybridization energy
V Matrix element of conduction band offset
W Matrix element of valence band offset
x Magnetic cation fractional content
α s-d exchange integral
β p-d exchange integral
ηq Fourier transform of magnetization
φ(z), ψ(r) Envelope functions

List of Abbreviations

1D, 2D, 3D One-, two-, three-dimensional


AF Antiferromagnetic
AP Atom probe
AFM Atomic force microscopy
AHE Anomalous Hall effect
AMR Anisotropic magnetoresistance
BMP Bound magnetic polaron
CMR Colossal magnetoresistance
CMS Condensed magnetic semiconductor
CPA Coherent potential approximation
DFT Density functional theory
DMS Dilute magnetic semiconductor
DFS Dilute ferromagnetic semiconductor
DMM Dilute magnetic material
DOS Density of states
EPR Electron paramagnetic resonance
FC Field cooled
fcc Face-centered cubic
FM Ferromagnetic
FMR Ferromagnetic resonance
GMR Giant magnetoresistance
HR High resolution
LDA Local density approximation
LSDA Local spin-density approximation
LT Low temperature
MBE Molecular beam epitaxy
MIT Metal-insulator transition
MFA Mean-field approximation
MOVPE Metal organic vapor phase epitaxy
(the same as MOCVD)
MCD Magnetic circular dichroism
MR Magnetoresistance
NC Nanocrystal
NN Nearest neighbor
NNN Next nearest neighbor
RE Rare earth
926 A. Bonanni et al.

RKKY Ruderman-Kittel-Kasuya-Yosida
SQUID Superconducting quantum interference device
TEM Transmission electron microscopy
TM Transition metal
TMR Tunneling magnetoresistance
VCA Virtual-crystal approximation
wz Wurtzite
XANES x-ray absorption near-edge structure
XAS x-ray absorption spectroscopy
XES x-ray emission spectroscopy
XRD x-ray diffraction
ZFC Zero-field cooled
zb Zinc-blende

Introduction

Dilute localized magnetic moments are usually associated with open d or f shells
of magnetic dopants but sometimes originate from nonmagnetic impurities and/or
lattice defects. The dilute systems attract considerable attention for a number of
reasons.
Firstly, even a minute concentration of uncoupled magnetic impurities can
change dramatically the transport and the optical characteristics. A relevant example
is the Kondo effect in metals ( Chap. 2, “Magnetic Exchange Interactions”)
used for the fabrication of efficient thermocouples at low temperatures. Another
is represented by intra-impurity optical excitations in semiconductors and oxides,
which dominate the optical properties in the spectral region below the fundamental
gap ( Chap. 3, “Anisotropy and Crystal Field”), a property exploited successfully
in stained glasses and broadband lasers [1].
Secondly, dilute systems bridge the physics and functionalities of semiconduc-
tors and magnetic materials. The existence of strong spin-dependent interactions
between bands and open magnetic shells, coexisting with spin-orbit interactions,
leads to giant magnetotransport and magnetooptical phenomena and opens the door
for spintronic functionalities, the control of magnetization and spin current by
light, strain, electric fields, and currents in both topologically trivial and nontrivial
systems.
Thirdly, the randomness in the distribution of interacting localized spins accounts
for novel magnetic properties, not encountered in chemically ordered crystals. More
specifically, dilute magnetic materials have gained the status of model systems for
assessing the nature of thermodynamic phase transformations in the presence of
spatial disorder, competing interactions, and frustration. Another appealing question
concerns quantum phase transformations in disordered systems and how localized
magnetism evolves into an itinerant or nonmagnetic ground state as a function of
the magnetic ion concentration.
Fourthly, materials science of dilute magnetic systems is surprisingly rich. It has
been increasingly clear that information about the nominal chemical composition is
rarely sufficient to account for the effective characteristics of these media. Actually,
to a large extent, their properties, including the nature of the magnetic ground
18 Dilute Magnetic Materials 927

state, depend on the way a given sample has been obtained. In particular, the
distribution of magnetic ions over the specimen volume (random vs nonuniform),
their incorporation into the lattice sites (substitutional vs interstitial), the resulting
spin configuration, and spin-spin interactions depend on the substrate employed,
growth conditions, co-doping, post-growth processing, surface reactions, and con-
tamination.
In this context the extensive use of nanocharacterization tools, such as syn-
chrotron radiation, microscopy, particle beam, and scanning probe methods, is
essential [2], as elaborated in  Chaps. 25, “Magnetic Scattering,” and  24,
“Magnetic Imaging and Microscopy.” Together with more traditional techniques
of magnetic resonance ( Chap. 26, “Electron Paramagnetic and Ferromagnetic
Resonance”) and optical magnetospectroscopy ( Chap. 10, “Magneto-optics and
Laser-Induced Dynamics of Metallic Thin Films”), these tools not only visualize
the micromagnetic characteristics but also reveal the distribution of magnetization
and magnetic constituents at the nanoscale as well as determine properties of energy
states relevant to magnetism with ever-improving spatial, chemical, spin, spectral,
and more and more often time resolution. Results of such combined structural
and spectroscopic investigations are now broadly employed to substantiate the
interpretation of experimental findings and to benchmark theoretical modeling,
particularly in newly developed systems [3].
Another specific feature of dilute magnetic materials is the small magnitude
of the magnetization, which makes these systems sensitive to both contamination
by ferromagnetic (FM) nanoparticles [4, 5, 6, 7] and experimental artifacts in
SQUID measurements [8]. A related timely question concerns the origin of a
temperature-independent anhysteretic magnetic signal often observed in nominally
diamagnetic solids. The search for new physics, such as contribution from orbital
currents, is now accompanied by an extensive use of spatially resolved analytic
tools, examination of samples from a range of suppliers and processed in various
ways [9].
Considering different hosts and magnetic ordering, dilute magnetic materials can
be grouped into six classes, as outlined in section “Dilute Magnetic Materials”.
A more detailed discussion of various aspects of dilute systems, together with
examples of experimental and theoretical findings, is presented in sections “Energy
States of Magnetic Dopants in Solids”, “Exchange Interactions Between Band
and Localized Spins”, “Effects of sp-d(f ) Exchange Interactions”, “Dominant
Spin-Spin Interactions”, and “Magnetic Properties of Dilute Magnetic Materials”.
The key parameter characterizing dilute magnetic systems is the concentration
of magnetic ions x. We warn the readers that various conventions exist in literature.
Here, by x we understand the fractional cation content. For instance, x in the
case of Mn-doped cadmium arsenide is defined by (Cd1−x Mnx )3 As2 . Within this
√ dopants per√unit volume is xN0 , where the
definition, the concentration of magnetic
cation concentration N0 = 4/a03 , 4/( 3a 2 c), 12/( 3a 2 c), and 8/a03 for the zinc-
blende and rock-salt, wurtzite, tetradymite, and elemental diamond-structure dilute
magnetic semiconductors, respectively.
Another convention of interest concerns the description of exchange coupling.
Here, the exchange interaction between a pair of localized spins is expressed in the
928 A. Bonanni et al.

form Hij = −2Jij S i · S j , whereas the exchange interaction between a band carrier
and a localized spin is expressed as Hi = −Jsp−d(f ) s · S i .

Dilute Magnetic Materials

It is natural to distinguish the six distinct families of dilute magnetic materials as


follows.

Dilute Magnetic Metals (DMMs)

In the limit of low concentrations of transition-metal impurities in nonmagnetic


metals, some dopants may acquire giant magnetic moments, or, conversely, impurity
moments can be screened by a cloud of band electrons (the Kondo effect), as
described in  Chaps. 14, “Magnetism of the Elements” and  2, “Magnetic
Exchange Interactions,” respectively. For higher concentrations, the energy scale
becomes determined by long-range carrier-mediated RKKY interactions between
localized spins ( Chap. 2, “Magnetic Exchange Interactions”). The randomness
of impurity distribution (quenched disorder) and frustration ( Chap. 2, “Magnetic
Exchange Interactions”) associated with alternations between FM and antiferromag-
netic (AF) coupling as a function of the spin-spin distance results in a specific
behavior, referred to as spin-glass freezing (section “Spin-Glass Systems”). Spin-
glass characteristics have been observed in numerous dilute metallic (crystalline
or amorphous) but also in nonmetallic systems [10]. Unusual properties (such as
a magnetic phase transition without anomalies in specific heat but with glassy-
like dynamics) have challenged standard theoretical models of statistical physics
and triggered the development of new conceptual and computational tools [11, 12]
impacting various fields of science [13] and computational hardware [14]. The
prevailing view is that in the case of long-range RKKY interactions specific to
DMMs, glassy dynamics reflect the wandering of spin configurations between
local free energy minima (hierarchical dynamics) [15], whereas the spontaneous
formation and annihilation of compact droplet excitations [16] describes better
nonmetallic materials such as DMSs, in which spin-spin interactions are short-
ranged. As in the case of other materials, “intrinsic” properties of DMMs are
affected by materials issues such as oxidation [17] and a nonuniform distribution of
the magnetic constituent [18].

p-Type Dilute Ferromagnetic Semiconductors (DFSs)

In these systems, a high density of delocalized or weakly localized holes leads to


long-range FM interactions between TM cations, which dominates over the superex-
change ( Chap. 2, “Magnetic Exchange Interactions”), and is well described by the
p–d Zener model [19] (section “p–d Zener Model”). The flagship example here
is (Ga,Mn)As [20], but to this category belong also other families of compounds
doped with transition metals, in which holes originate from point defects, e.g.,
18 Dilute Magnetic Materials 929

(Pb,Sn,Mn)Te [21], or acceptor impurities, e.g., (Cd,Mn)Te/(Cd,Mg)Te:N [22],


rather than from Mn. Groundbreaking spintronic functionalities demonstrated for
these systems [23,24], rely on the strong p–d coupling between localized spins and
hole carriers as well as on a sizable Luttinger spin-orbit interaction in p-like orbitals
forming the valence band or originating from inversion asymmetry of the structure.
The reported magnitudes of Curie temperature TC reach 200 K in (Ga,Mn)As
[25, 26, 27], (Ge,Mn)Te [28, 29], and (K,Ba)(Zn,Mn)2 As2 [30] with less than 10%
of Mn cations, xeff < 0.1, measured by saturation magnetization in moderate
fields, μ0 H  5 T (section “p-Type Dilute Ferromagnetic Semiconductors”).
A numerous functionalities evident in (Ga,Mn)As and related DFSs (magnetiza-
tion control by an electric field, current-induced domain-wall motion, anisotropic
tunneling magnetoresistance) are now transferred to magnetic metals supporting
ferromagnetism above room temperature. A specific feature of p-type DFSs is the
interplay of carrier-mediated ferromagnetism with carrier localization, which results
in spatial fluctuations of magnetization and superparamagnetic signatures generated
by nonuniformities in carrier density that appear in the vicinity of the metal-to-
insulator transition [31, 32].

Other Dilute Ferromagnetic Semiconductors

The Goodenough-Kanamori-Anderson rules ( Chap. 2, “Magnetic Exchange


Interactions”) indicate in which cases a superexchange can lead to FM short-
range coupling between localized spins. According to experimental and theoretical
studies, such a mechanism operates for Mn3+ ions in GaN and accounts for TC
values reaching about 13 K at x ≈ 10% in semi-insulating wurtzite Ga1−x Mnx N
[33] (section “Dilute Ferromagnetic Insulators and Topological Insulators”). The
dependence of TC on x corroborates the percolation theory [34] that implies
the coexistence of percolating and non-percolating FM clusters at T > 0.
Ferromagnetic superexchange and its variants dominate in magnetic semiconductors
[35] such as Cr-spinels and CdCr2 Se4 [36] (TC = 130 K) as well as in
EuO (TC = 68 K) and EuS (TC = 16 K) [37], known for showing colossal
magnetoresistance effects in the vicinity of TC [38]. By electron doping (oxygen
vacancies, Eu substituted by Gd), the TC of Eu compounds can be enhanced by
about 50 K in agreement with the RKKY theory that explains also TC ≈ 160 mK in
n-Zn1−x Mnx O:Al with x = 0.03 [39].

Dilute Magnetic Semiconductors (DMSs)

This material family was initially named semimagnetic semiconductors [40], as


it comprises standard semiconductors doped with magnetic impurities that are
randomly distributed, electrically inactive, and do not undergo any long-range spin
ordering [41,42]. Typical representatives are (Cd,Mn)Te, (Zn,Co)O, (Hg,Fe)Se, and
(Pb,Eu)S. Short-range AF superexchange between transition metal ions or long-
range dipolar interactions between rare-earth magnetic moments lead to spin-glass
930 A. Bonanni et al.

freezing at Tf < 1 K for x < 0.1 (section “Spin-Glass Systems”). The use of magne-
tooptical and quantum transport techniques has revealed and quantified the influence
of sp–d coupling upon the exciton and Landau-level spectra, quantum Hall effect,
quantum localization, universal conductance fluctuations, and spin dynamics in
DMS systems of various dimensionality [43] (section “Effects of sp-d(f ) Exchange
Interactions”) as well as made it possible to demonstrate electrical spin injection
[44]. Conversely, the spin subsystem has been probed by the Faraday effect [45]
or quantum noise [46] (section “Spin-Glass Systems”). The physics of bound
magnetic polarons, a single electron interacting with spins localized within the
confining potential of impurities or quantum dots, has been advanced in DMSs
[47] (section “Bound Magnetic Polarons”). Most of the end-member compounds,
for instance, MnSe and EuTe, are antiferromagnets, which together with related
materials, like CuMnAs, constitute the building blocks for the emerging field of AF
spintronics [48]. Another ultimate limit of DMSs is represented by qubit systems
consisting of single magnetic ions in single quantum dots [49].

Dilute Magnetic Topological Materials

In these systems magnetization-induced giant p–d exchange spin-splitting of


Dirac cones turns helical states into chiral states. Striking consequences of this
transformation include (i) the precise quantization of Hall conductance σxy = e2 / h
demonstrated for thin layers of FM (Bi,Sb,Cr)2 Te3 at millikelvin temperatures [50],
as predicted theoretically [51]; (ii) the efficient magnetization switching by spin-
locked electric currents in these ferromagnets [52]; and (iii) the formation of Weyl
semimetals from 3D Dirac spin-glass materials, such as (Cd,Mn)3 As2 or strained
(Hg,Mn)Te [53] (section “Dilute Ferromagnetic Insulators and Topological Insula-
tors”). Furthermore, inverted band ordering specific to topological matter enhances
the role of the long-range interband Bloembergen-Rowland contribution to spin-spin
interactions, resulting in higher spin-glass freezing temperatures Tf in topological
semimetals, such as (Cd,Mn)3 As2 and (Hg,Mn)Te, compared to topologically trivial
II-VI Mn-based DMSs [54]. This exchange, taken into consideration within the p–d
Zener model [55] (section “p–d Zener Model”) and named Van Vleck’s mechanism,
was proposed to lead to ferromagnetism in topological insulators of V-, Cr-, or Fe-
containing bismuth/antimony chalcogenides [51]. However, according to ab initio
studies and corresponding experimental data, the superexchange in the case of V and
Cr and the RKKY coupling in Mn-doped films appear to contribute significantly to
ferromagnetism in tetradymite V2 -VI3 compounds [56]. Ferromagnetism mediated
by carriers in surface or edge Dirac cones is rather weak for realistic values of spatial
extension of these states into the bulk [23].

Heterogenous Magnetic Semiconductors and Oxides

These materials are characterized by a highly nonrandom distribution of magnetic


elements brought about by binodal or spinodal nanodecomposition (chemical
phase separation) or nanoprecipitation (crystallographic phase separation) [3]
18 Dilute Magnetic Materials 931

(section “Heterogenous Magnetic Semiconductors and Oxides”). Here, nanoregions


with high concentrations of magnetic cations, denoted as condensed magnetic
semiconductors – CMS [57] – account for superparamagnetic, superferromagnetic
( Chap. 20, “Magnetic Nanoparticles”), or FM-like properties persisting typically
above room temperature. The CMSs appear in the shapes of nanodots (the dairiseki
phase) or nanocolumns (the konbu phase) buried in semiconducting or insulating
hosts [58]. Such nanocomposites, fabricated in a self-organized fashion and
containing metallic magnetic nanostructures, have a potential for applications in
electronics, spintronics, photovoltaics, plasmonics, and thermoelectrics [3].

Energy States of Magnetic Dopants in Solids

A good starting point for the description of dilute magnetic materials is the
Vonsovskii model, according to which the electron states can be divided into two
categories: (i) localized magnetic states, typically originating from open d or f
atomic shells, and (ii) extended band states built up of s, p, and also d atomic
orbitals. The understanding of how particular hosts perturb open shells of magnetic
impurities or defects is built on two pillars:

• Crystal-field theory: This is essentially a first-order perturbation theory that


shows how electrostatic potentials generated by nonmagnetic ligands split open
multi-electron atomic shells of magnetic dopants ( Chap. 3, “Anisotropy and
Crystal Field”) or point defects depending on their location in the lattice.
These splittings, and thus the resulting charge and spin configurations of the
impurity ground and excited states, are described for various crystal structures
by the Tanabe-Sugano diagrams as a function of the crystal-field strength
( Chap. 4, “Electronic Structure: Metals and Insulators”). In more general
terms, characteristics of possible states, including the forms of corresponding
spin Hamiltonians [59], are dictated by the spin-containing group theory, whose
parameters are provided experimentally, and reproduced increasingly success-
fully by relativistic ab initio computations ( Chap. 3, “Anisotropy and Crystal
Field”). In general, electronic energies are lowered by coupling to the lattice,
by a local deformation. This is particularly relevant for orbitally degenerate
states, the mechanism described in terms of a static or dynamic Jahn-Teller effect
( Chap. 3, “Anisotropy and Crystal Field”). In extreme cases coupling to the
lattice can lead to a negative value of the intra-site correlation energy Ueff .
• Anderson Hamiltonian: The Anderson model of magnetic impurities ( Chap. 4,
“Electronic Structure: Metals and Insulators”) describes a competition between
the intra-atomic Coulomb energy U (that accounts for the presence of open d
and f shells) and quantum hopping of electrons between these localized shells
and extended band states (quantified by the hybridization matrix element Vkd ).
Thus, hybridization enlarges the localization radius of magnetic electrons, which
lowers their intra-site correlation energy. This effect explains (i) the existence
of numerous levels corresponding to subsequent charge states of impurities or
defects within the semiconductor energy gap (the Haldane-Anderson mecha-
932 A. Bonanni et al.

nism) [60] and (ii) the disappearance of localized magnetic moments for Vkd >
U, the case of many combinations of magnetic impurities and metallic hosts.
However, as long as localized magnetic moments survive in a given host, so that
the corresponding d states do not contribute to the Fermi volume, the presence of
hybridization does not invalidate the expectations of the crystal-field theory but
only modifies the magnitudes and microscopic origins of particular crystal-field
parameters.

Figure 1 depicts ground-state energy levels corresponding to two charge states of


cation-substitutional TM ions in zinc-blende or wurtzite II-VI and III-V compounds.
These two states have the donor and acceptor character (lower and upper Hubbard
bands, respectively), and the energy difference between them is the effective
correlation energy Ueff . For either of these material families, the energies of TM
levels are to a good approximation universal, in the sense that they can be presented
in one plot if the relative positions of the band edges in particular compounds
are shifted according to the band offsets known from heterostructure studies.
For particular charge states, the spin localized on the TM ion assumes in these
compounds the highest possible value, according to Hund’s rule. In the case of ions
with the orbital momentum L = 0, the Jahn-Teller effect and spin-orbit coupling
affect positions and splittings of the levels. If there is no net magnetic moment
associated with the ion, the case of Fe2+ in II-VI compounds, magnetization is of
the Van Vleck type at low temperatures.
The diagrams in Fig. 1 make it possible to assess the electrical activity of a given
TM impurity in a given host and the variation of its charge state with the extrinsic
doping. In particular, the levels residing within the bandgap act as carrier traps and,
thus, may serve for obtaining a semi-insulating material [62]. Alternatively, TM
impurities donate electrons if the corresponding donor state is above the bottom of
the conduction band (e.g., Sc in CdSe) or holes if the acceptor state is below the top
of the valence band (e.g., Mn in GaAs). It is important to note that these resonant
states are accompanied by shallow donor or acceptor levels, respectively, created by
the Coulomb potential of the charged TM impurities, i.e., Sc3+ in CdSe and Mn2+
in GaAs.
Another type of level that is not depicted in Fig. 1 is the so-called charge transfer
state or Zhang-Rice polaron [63]. It appears if the magnitude of p–d hybridization
is sufficiently large (the strong coupling limit) to result in bound states of holes on
anions surrounding isoelectronic (neutral) TM impurities, the case of, e.g., Fe3+
in GaN and Mn2+ in ZnO [64]. The same mechanism accounts for a gradual
increase of the Mn2+ acceptor binding energy in III–V compounds on going from
the antimonides, through arsenides and phosphides, to the nitrides, as a contribution
of p-d hybridization to the binding energy increases when the cation-anion distance
diminishes [23].
A similar set of data has been collected for rare-earth impurities in various
families of compounds (see Ref. [42] for the case of II–VI semiconductors). On
the other hand, no such data are so far available for newly developed topological
materials, such as thallium/nobium pnictides and bismuth/antimony chalcogenides
in which TM doping plays an increasingly important role.
18 Dilute Magnetic Materials 933

ZnTe
ZnS
ZnSe
A(0/-)

CdTe
CdS
ZnO

CdSe

HgTe
2
Energy (eV)

D(0/+)

Sc Ti V Cr Mn Fe Co Ni Cu
0 1 2 3 4 5 6 7 8 9
d d d d d d d d d

-2

AlN

AlSb
AlAs
AlP
GaP

Ti V Cr Mn Fe Co
GaSb
GaAs

1 2 3 4 5 6
d d d d d d
GaN

2
InP

InSb

A(0/-)
Energy (eV)

InAs
InN

0 D(0/+)

-2

Fig. 1 Approximate positions of ground-state levels derived from d shells of cation-substitutional


transition-metal impurities in two different charge states relative to the conduction and valence
band edges of II-VI (upper panel) and III-V (lower panel) compounds. Triangles and squares
denote d N /d N −1 donor (D) state (donating the electron, N → N − 1, becomes positively
charged) and d N /d N +1 acceptor (A) states (accepting the electron, N → N + 1, becomes
negatively charged), respectively. Typically, the wave function of these states contains a substantial
contribution from p orbitals of neighboring anions. These diagrams include neither shallow donor
or acceptor levels accompanying resonant states in the conduction and valence band, respectively,
nor mid-gap hole traps appearing for strong p–d hybridization, the case of oxides and nirides.
(Adapted from Langer et al. [61] and Zunger [60] updating the values of band offsets)
934 A. Bonanni et al.

Detailed information on energy levels corresponding to the ground and excited


states at a given charge/spin configuration and coupling to the lattice is essential for
designing light emitters, particularly broadband optically pump lasers of TM-doped
oxides or chalcogenides [1]. If an impurity level participating in optical transitions
overlaps with a band continuum, the absorption spectrum may assume a form of
the Fano resonance. In metals, in turn, resonant scattering of carriers is expected
if the Fermi energy εF coincides with the ground-state level of impurities, the case
presented in Fig. 2a. However, if the Fermi level is located in-between impurity
Hubbards bands, electron scattering rate is enlarged, too – this time by the presence
of the Abrikosov-Shul resonance at the Fermi level (Fig. 2b), which is formed by
electrons dynamically bound to localized spins, according to theory of the Kondo
effect in the strong coupling limit ( Chap. 2, “Magnetic Exchange Interactions”).
A number of phenomena appear when increasing the content of the magnetic
constituent. Below, we consider first insulator-to-metal transitions transforming
localized centers into extended states, either merging with the existing bands or
forming additional bands. Next, the case of resonant levels is described. Finally,
the question is addressed what happens if the impurity concentration surpasses the
solubility limit. The origin and effects of exchange coupling between carriers and
localized spins as well as the mechanisms and consequences of interactions among
localized spins are discussed in subsequent sections.
Anderson-Mott transition: Within this model, localization of carriers in their
parent band results from quantum interference of many-body and single-carrier
scattering amplitudes [65]. For disorder-modified hole-hole interactions and hole

E (a) E (b)

HF HF

DOS DOS
E (c) E (d)

HF

HF

DOS DOS

Fig. 2 Schematic forms of density of states (DOS) if magnetic impurities introduce states
overlapping with a continuum of band states; (a) resonant state; (b) Abrikosov-Shul resonance
at the Fermi level εF appearing below the Kondo temperature; (c) Efros-Shklovskii Coulomb gap
at εF in the impurity DOS appearing if inhomogeneous broadening of the impurity band is larger
than lifetime broadening; (d) hybridization gap for Kondo insulators
18 Dilute Magnetic Materials 935

scattering by ionized cation-substitutional Mn2+ acceptors in III-V semiconductors,


the critical concentration pc is about 1016 cm−3 for Insb, 1020 cm−3 for GaAs, and
above 4 × 1021 cm−3 for GaN. These values of pc are consistent with the empirical
Mott criterion pc ≈ (0.26/aB )3 , where aB is an effective Bohr radius of the dopant
in question, here the hole localization radius in a single neutral Mn3+ = Mn2+ + h
acceptor center [23]. Obviously, Anderson-Mott physics is relevant to magnetic
semiconductors also in the case when carriers originate from extrinsic dopants, such
as oxygen vacancies in EuO or In in (Cd,Mn)Se. It should be noted that no available
ab initio method can describe the Anderson-Mott localization.
Mott-Hubbard transition: Zinc-blende MnTe appears to be a localized antiferro-
magnet meaning that 3d levels do not undergo a Mott-Hubbard insulator-to-metal
transition (see  Chap. 17, “Magnetic Oxides and other Compounds”). In contrast,
zinc-blende MnAs is presumably an itinerant ferromagnet, which indicates that at
a certain value of Mn density in (Ga,Mn)As delocalization of Mn 3d levels occurs.
This issue, despite its relevance to a number of dilute magnetic semiconductors,
has not yet been experimentally explored, presumably because materials with low
concentrations of magnetic impurities are usually investigated. Theoretical studies
on disorder effects in the Mott-Hubbard transition are under way [66].
Coulomb gap and charge ordering: Inter-site interactions within the impurity
band give rise to the Efros-Shklovskii Coulomb gap in the one-carrier density of
states at the Fermi level [67], as shown in Fig. 2c. Let’s consider a concentration
of magnetic impurities low enough that no Mott-Hubbard transition occurs but
sufficiently large for the band carriers to be delocalized and the Fermi level pinned
by the resonant state. Under these conditions, the Coulomb gap in the impurity band
decouples it from the continuum and precludes resonant scattering of band carriers.
At the same time, the momentum relaxation rate for scattering by ionized impurities
is diminished, as the inter-site Coulomb interactions impose spatial ordering of
impurity charges. Accordingly, record-high electron mobilities up to 2×107 cm2 /Vs
were observed in zero-gap semiconductors when the Fermi level was pinned by
either resonant donors (HgSe:Fe [68]) or acceptors [(Hg,Mn)Te:Cu [69]]. A similar
effect was found in (Sn,Gd)Te [70].
Hybridization gap and Kondo insulators: The Kondo effect vanishes if the
magnitude of RKKY coupling between magnetic impurities surpasses the Kondo
temperature, which for TM impurities in metals occurs at densities typically below
1%. However, there is a class of compounds including SmB6 , Ce3 Bi4 Pt3 , FeSi
[71] for which the Anderson Hamiltonian remains valid. For the periodic case in
question, there exists a solution that points to the presence of a gap in the density of
states at the crossing of the magnetic level with a continuum of band states, the case
illustrated in Fig. 2d. If the Fermi level is located in this gap, the bulk is insulating
at low temperature, the case for SmB6 that possesses additionally topologically
protected surface states conducting in the limit of zero temperature [71]. According
to one of the scenarios, in heavy-fermion metals ( Chap. 15, “Metallic Magnetic
Materials”), the Fermi level is located just above or below the hybridization gap,
where the density of states is largely enhanced by a sizable component coming from
magnetic shells that are periodically arranged but highly localized around parent
atoms. How dilution of the magnetic constituent by nonmagnetic atoms affects
936 A. Bonanni et al.

fragile coherence leading to the hybridization gap and in some cases to topological
states remains largely an open question.
Beyond the solubility limit: The stability of a particular alloy A1−x Bx against the
decomposition is described by the mixing energy, i.e., the alloy energy in respect to
the weighted energies of the end compounds:


E(x) = EAB − (1 − x)EA − xEB . (1)

Due to the aforementioned hybridization, d shells perturb bonding, which often


makes
E(x) convex, meaning that the alloy is unstable against binodal decompo-
sition into the end compounds. In reality the degree of nanoscale nonuniformities in
the alloy composition and in the local crystal structure, which set at a given growth
or annealing temperature, are determined by complex interplay between energy,
entropy, and diffusion barriers in the material volume and at the growth surface.
These effects are known in the DMS literature as spinodal nanodecomposition, and
their dynamics is described by the Cahn-Hilliard equation [3]. As a result of the
spinodal nanodecomposition, electronic structure of magnetic dopants may vary on
the nanoscale and depend more on growth and processing procedures than on an
average chemical composition, as elaborated in section “Heterogenous Magnetic
Semiconductors and Oxides”.

Exchange Interactions Between Band and Localized Spins

The key feature of dilute magnetic materials is the existence of strong spin-
dependent coupling between magnetic ions and band carriers, which accounts for
the unique low-energy properties – transport, optical, and thermodynamic char-
acteristics of these systems. Neglecting non-scalar corrections that can appear for
magnetic ions with nonzero orbital momentum, L = 0, a one-electron Hamiltonian
that perturbs dynamics of the Bloch electrons assumes the Kondo form, HK =

i Hi , where
K

HiK = [V (r − R i ) − J (r − R i )s · S i ]. (2)

Here V (r − R i ) is a spin-independent potential change introduced by the magnetic


impurity, and J (r − R i ) is a short-range exchange energy operator between the
carrier spin s and the localized spin S residing at R i . In general, V (r − R i ) contains
a long-range Coulomb part, present if the magnetic impurity is not isoelectronic,
and a short-range alloy potential Vs (r − R i ), giving rise to a central-cell correction
or a band offset in the impurity and alloy nomenclature, respectively.
There are two mechanisms contributing to the spin-dependent part of the Kondo
coupling J : (i) the exchange part of the Coulomb interaction between the effective
mass and localized electrons, potential exchange which is FM, J > 0, and (ii)
quantum hopping (hybridization) between band and local states resulting in a spin-
dependent partial delocalization of electrons, which lowers their kinetic energy to
18 Dilute Magnetic Materials 937

Table 1 Magnitudes of the Mn+ meV Eu+ meV


intra-atomic exchange
4s-3d 392 6s-4f 52
energies J for singly ionized
free atoms Mn+ and Eu+ , as 4p-3d 196 6p-4f 33
collected by Dietl et al. [72] 4d-3d 107 5d-4f 215

a degree dependent on the mutual spin orientation. This mechanism, referred to as


the kinetic exchange, can be of either sign. It also contributes significantly to the
magnitude of Vs .
Potential exchange: Since the exchange part of the Coulomb interaction is short-
ranged, the potential exchange for the cation-substitutional magnetic impurities is
sizable only if the carrier wave is appreciable at the cation sites. Table 1 contains
exchange energies deduced from the energy differences ES+1/2 − ES−1/2 for an
electron occupying various outer shells in free standing Mn+ (S = 5/2) and
Eu+ (S = 7/2) ions. In the picture of ionic bonding, the value J4s−3d = 0.39 eV
would determine entirely the exchange energy Js−d in the s-type band, as there
is no hybridization between the Γ6 states and the levels derived from the d shell
(eg and t2 states). Actually, the covalency (s–s hybridization) leads to spreading
of the electron wave function between cation and anion sites. In accord with this
insight, the values of Js−d (named N0 α in the DMS literature), as determined from
magnetooptics experiments, are reduced to the range 190  N0 α  320 meV for
a series of paramagnetic II-VI compounds doped with Mn as well as with Cr, Fe,
and Co [73], in agreement with ab initio computations [74]. This reduction, as seen
by observations of the spin-splitting, is even stronger in the case of Mn-based III-V
DMSs, in which Mn ions are negatively charged [75].
In the case of Eu chalcogenides and mono-oxide, the magnitude of spin-splitting
of the conduction band [76] points to the exchange energy J = 0.17 eV. This value
is consistent with the data for Eu+ displayed in Table 1, as the wave function of
conduction band electrons is primarily built of Eu 5d atomic orbitals.
Kinetic exchange: By employing a canonical transformation, Schrieffer and
Wolff demonstrated how to derive the kinetic part of the Kondo Hamiltonian from
the Anderson Hamiltonian for impurities with a single open orbital embedded in a
nonmagnetic metal. Since then a number of generalizations have been proposed,
in particular taking into account realistic level structures of particular magnetic
impurities but assuming that the carrier wave functions are in the form of plane
waves [77]. The kinetic term overweighs the potential exchange and accounts for
the AF sign of J leading to the Kondo effect in diluted magnetic metals ( Chap. 2,
“Magnetic Exchange Interactions”).
In the case of DMSs, hybridization between periodic parts uk of the Bloch
wave functions and magnetic shells is considered. In particular, there is a strong
hybridization between Γ8 and t2 states, which affects their relative position and leads
to large values of |Jp−d ≡ N0 β| ≈ 1 eV collected for a number of tetrahedrally
coordinated Mn-based DMSs in Fig. 3. If the relevant effective mass state is above
the t2g level (the case of Mn-based DMSs), then β < 0, but otherwise β can be
938 A. Bonanni et al.

Exchange energy, –N0β (eV)


3 β = – 0.054 eV nm3 ZnO

2 CdS
CdSe ZnSe GaN
ZnTe ZnS
1 CdTe GaAs
InAs

0
ZnO
-1 solid symbols: photoemission, XAS GaN
open symbols: magnetooptics

-2
0 10 20 30 40 50

Cation concentration, N0 (nm-3)

Fig. 3 Compilation of experimentally determined energies of the p-d exchange interaction −N0 β
for various Mn-based DMSs as a function of the cation concentration N0 . Solid symbols denote
the values evaluated from photoemission and x-ray absorption spectra. The values shown by
open symbols were determined within the virtual-crystal and mean-field approximations (i.e.,
neglecting effects of strong coupling) from excitonic or band splittings in a magnetic field. Solid
line corresponds to a constant value of β across the DMS series. (The data collected by Dietl and
Ohno [23])

positive [the case of Zn1−x Crx Se [78]]. These findings are supported by theoretical
considerations [73, 74].
It should be emphasized that the role of hybridization depends crucially on the
symmetry of band states and the parity of the localized state. The description of
carrier-impurity interactions had to be reconsidered entirely in rock-salt DMSs, such
as (Pb,Mn)Te [72], as well as when rare-earth atoms are involved, as in (Pb,Eu)Se
[72].
Finally, we note that appropriately far away from the band extremum, the kp
method [79] ceases to work and, additionally, the term exp(ikr) in the carrier wave
function generates additional hybridization with the open magnetic shells.

Effects of sp-d(f ) Exchange Interactions

Spin-Splitting of Extended States: Weak Coupling

Within the time-honored virtual-crystal approximation (VCA) and mean-field


approximation (MFA), the host crystal symmetry is restored by replacing HK
(Eq. 2) by its translationally invariant form. Taking into account only the short-range
part in the spin-independent term,
18 Dilute Magnetic Materials 939


Hav = x[Vs (r − R j ) − J (r − R j )s · S], (3)
j

where now j runs over all atoms in the sublattice containing magnetic ions (typically
the cation sublattice) and S is a thermodynamic average value of spin operators
Si . In terms of S, macroscopic spin magnetization M o (T , H ) = −g0 μB xN0 S,
where g0 = 2.0. In the case of impurities with nonzero orbital momentum L (e.g.,
the rare-earth case [80]), M o (T , H ) = −gJ μB xN0 J , and S = (gJ − 1)J 
( Chap. 3, “Anisotropy and Crystal Field”).
The above approach is suitable for describing the influence of magnetic impu-
rities onto the spin-resolved band structure under two conditions: (i) there is no
correlation in the distribution of magnetic ions, as the presence of any superstructure
would lower point symmetry; (ii) the total potential introduced by a single impurity
is too weak to give rise to bound states, so that the effect of HK on extended states
can be treated perturbatively (weak coupling limit).
Within a first-order perturbation theory and in the spirit of the multiband
effective mass approximation [79, 41], the effect of Hav is described by matrix
elements ui |Vs |ui  and ui |J |ui , where ui are periodic Kohn-Luttinger amplitudes
corresponding to bands at the relevant band extremum. In the case of substitutional
magnetic impurities in zinc-blende DMSs, for which typically bands with Γ6 , Γ7 ,
and Γ8 symmetries at k = 0 should be taken into account, there are four relevant
matrix elements: V = uc |Vs |uc  and α = uc |J |uc ; W = uv |Vs |uv  and
β = uv |J |uv  involve s-type (uc ) and p-type (uv ) periodic parts of the Bloch
wave functions at k = 0, respectively. This formalism can be extended into low-
dimensional nanostructures and topological materials, particularly by mapping the
multiband effective mass approach onto a multiorbital tight-binding approximation
[81, 82].
In general, due to kp coupling between bands and the spin-orbit interaction, the
impurity-induced shift and spin-splitting of bands depend in a complex way on k,
M 0 , and the angle between them. However, at k = 0 this shift is xN0 V and xN0 W
for the Γ6 and Γ8 band, respectively, whereas the corresponding splitting is h̄ωs =
αMo /gμB for the Γ6 band and h̄ωs(h) = βMo /gμB for the heavy-hole Γ8 band at
k = 0. For x = 0.05, N0 β = −1 eV (see Fig. 3), and saturated spins S = 5/2,
(h)
|h̄ωs | = 0.125 eV.
A combination of the MFA and the effective mass formalism has been success-
fully applied to describe spin-splittings of bands and excitonic spectra in bulk,
films, and various dimensionality quantum structures of DMSs as well as spin-
splittings of Landau levels in these systems as a function of temperature and a
magnetic field [41, 42, 43, 83]. Such studies have provided quantitative information
on the magnitudes of sp–d(f ) exchange integrals in various families of DMSs;
some examples are collected in Fig. 3. The VCA and MFA formalism can readily
be employed to assess spin-splittings of boundary states specific to topological
materials [84, 51].
When the relevant physics involves length scales greater than an average distance
between magnetic ions (see  Chap. 7, “Micromagnetism”), it is convenient to work
940 A. Bonanni et al.

within the continuous medium approximation. In the case under consideration, it


assumes a form of the molecular-field approximation, in which the spin-independent
part leads to a position-independent energy shift, whereas the spin-dependent part
is given by

Hsp−d(f ) = Jsp−d(f ) s · M(r)/N0 gμB , (4)

where Jsp−d(f ) is the relevant exchange energy (Js−d ≡ N0 α and Jp−d ≡ N0 β in


tetrahedrally coordinated semiconductors) and M(r) is the local spin magnetization
that shows spatial variations due to the presence of, for instance, magnetic domains
or thermodynamic fluctuations of magnetization.
According to Eq. 4, band carriers experience a molecular field H ∗ =
Jsp−d(f ) M/N0 gg ∗ μ2B , where g ∗ is the carrier Landé factor. Similarly, the spins
are in a molecular field generated by band carriers, H s = Jsp−d(f ) nP /N0 gμB ,
where n and P are carrier density and spin polarization, respectively. The field H s
displaces the EPR frequency of localized spins (the Knight shift), which constitutes
another way of determining the magnitude and sign of sp–d(f ) exchange energies
[70]. This method is particularly useful if spin-orbit interactions make g ∗ and g
different, ensuring the absence of a spin bottleneck [85].
Exchange spin-splitting of bands and associated spin polarization of carriers
allows one to generate spin currents essential for spin injection, spin-transfer torque,
GMR, TMR, and other spintronic capabilities of magnetic and dilute magnetic
systems [23, 24].

Spin-Splitting of Extended States: Strong Coupling

If the local perturbation of the crystal potential introduced by a single impurity HiK
is strong enough to bind a carrier even in the absence of a Coulomb potential, VCA
and MFA do not describe adequately the influence of magnetic doping on the band
states. This is particularly the case for nitrides and oxides, where a relatively short
bond length results in a sizable p–d hybridization. In this strong coupling case, the
spectral density of states Asz (ω) was determined by a non-perturbative generalized
alloy theory for noninteracting Heisenberg spins. A maximum of Asz (ω) provides
the position of the valence band top (if holes can be trapped) for particular spin
orientations in respect to magnetization Mz (T , H ) of localized spins at given values
of the integrals W and β as well as of the width of the potential well, which
controls the binding ability of the neutral magnetic impurity [86]. This approach
demonstrates that both the band shift and splitting can even assume an opposite sign
to that expected within the VCA and MFA, as extended states are pushed away from
the impurity, if the spectrum contains a bound state for a given spin orientation. As
shown in Fig. 3, the values of β determined from excitonic splittings neglecting this
effect appear as anomalous and do not follow chemical trends in the case of nitrides
and oxides.
18 Dilute Magnetic Materials 941

Another case is represented by magnetic impurities that are electrically active,


acting as donors or acceptors, so that HiK contains also a long-range Coulomb
potential. The effect of such impurities on spin-splitting of band states depends on
whether the dopants are occupied or ionized. In the former case, a band carrier is
a subject of exchange coupling not only to the spin of the magnetic ion but also to
the spin of the carrier bound to it [87]. If magnetic donors or acceptors are ionized,
the Coulomb potentials renormalize the probability of finding a band carrier in the
vicinity of magnetic ions [87, 75]. In either case, the band spin-splitting and the
apparent sp–d exchange integral can be substantially modified compared to cases
when magnetic ions are not electrically active.

Alloy and Spin-Disorder Scattering: Weak Coupling

In general, the Hamiltonian describing alloy and spin-disorder scattering Hsc is


determined by a difference between the full Kondo Hamiltonian (Eq. 2) and its VCA
and MFA version (Eq. 3) serving to determine the band structure of magnetic alloys.
In order to obtain the relevant scattering cross section, two principal approaches
have been put forward. The first is the phase-shift theory introduced by Friedel
as the d resonance model and then successfully applied to describe the residual
resistivity [88], electron spin-resonance linewidth [89], and the electronic structure
[90] of noble metals with transition-metal impurities.
Another approach makes use of a perturbation theory whose weak coupling
version corresponds to the Born approximation. A particularly transparent formula
for the momentum relaxation time τ is obtained assuming that kB T is greater than
energies of magnetic excitations as well as that the spin-splitting h̄ωs is much
smaller than εF ( Chap. 9, “Magnetotransport”). For cubic materials and s-type
wave functions and in terms of the band shift V and exchange integral α one obtains
for cubic materials and assuming that the z direction is along the magnetization
Mo (T , H ), a derivative of the de Gennes-Friedel and Béal-Monod-Weiner formulae
[91],

1/τ = 1/τal + 2/τsx + 1/τsz . (5)

The alloy and spin-disorder scattering rates are given by

1/τal = π xN0 V 2 ρ(ε)/h̄;


π α 2 ρ(ε)
1/τsx = k B T χ⊥ ;
4h̄g02 μ2B (6)
   
π α 2 ρ(ε) x(1 − x) dMo (T , H ) 2
1/τsz = k B T χ + ,
4h̄g02 μ2B N0 dx

where magnetic susceptibilities of localized spins


942 A. Bonanni et al.

χ⊥ = Mo (T , H )/H and χ = ∂Mo (T , H )/∂H, (7)

enter the formulae for τsi via the fluctuation-dissipation theorem. The term in
square brackets in Eq. 6 describes the effect of chemical disorder (inhomogeneous
broadening). For spherical bands and in the 3D case, the density of states (DOS)
ρ(ε) = m∗ k(ε)/π 2 h̄2 ; for a single electric subband in 2D systems, k(ε) should be
replaced by π/L̃W , where the participation length

L̃W = 1/ dz|φ(z)|4 . (8)

Here, φ(z) is the envelope function describing the carrier confinement, and the
integration extends over the region where the magnetic ions reside.
In the vicinity of the critical temperature for magnetic ordering, where χ tends
to diverge, several additional effects have to be considered in evaluating τsi (T ),
including the q dependence of χ , screening of magnetization fluctuations by carrier
redistribution, and finite quantum coherence length [92]. A careful description of
the pertinent effects will be essential in order to explain the behavior of resistance
and its temperature derivative at criticality [93].
Since the exchange part of Hsc breaks spin rotation symmetry, spin-disorder
scattering, together with spin-orbit effects (section “Interplay of sp–d(f ) Exchange
Interactions and Spin-Orbit Coupling”), accounts for carrier spin relaxation in
magnetic systems as well as controls the magnitude of quantum interference
effects underlying Anderson-Mott localization and mesoscopic transport phe-
nomena (section “Quantum Localization and Mesoscopic Phenomena: Colossal
Magnetoresistance”).

Alloy and Spin-Disorder Scattering: Strong Coupling

A manifestation of strong coupling in spin-disorder scattering is the Kondo enhance-


ment of the resistivity below a certain temperature, which appears for AF coupling
between a single localized impurity spin and spins of the carrier sea (Fig. 2b and
 Chap. 2, “Magnetic Exchange Interactions”). The effect has been widely observed
in diluted magnetic metals but not yet in degenerate p-type diluted magnetic
semiconductors, presumably because a relative small magnitude of DOS compared
to metals shifts the Kondo phase to rather low temperatures in DMSs, where it
competes with localization phenomena and spin-orbit effects.

Bound Magnetic Polarons

The concept of bound magnetic polarons (BMPs), spin complexes consisting of


a donor electron and a bath of localized impurity spins residing within electron’s
confinement region, was introduced by von Molnar and Kasuya in the context
18 Dilute Magnetic Materials 943

of europium chalcogenides. Thus, BMPs exist in the strongly localized regime


of the Anderson-Mott transition in magnetic and dilute magnetic semiconductors
containing carriers or excitons trapped in singly occupied localized states [94].
Since a typical number of localized spins contributing to a BMP is greater
than 10, they can be treated classically and described within the molecular-field
approximation, i.e., characterized by the order parameter M(r). The starting point
of the Dietl-Spałek model of BMPs [47] is the electron spin Hamiltonian s · 
with eigenvalues describing spin-split electron energies ± 12
. In this approach, the
exchange part of


α

s−d = | drM(r)|ψ(r)|2 |, (9)
g0 μB

where α is the s–d exchange integral and ψ(r) is the electron envelope function.
Hence, in the presence of a collinear magnetic field B and an average magnetization
M 0 of bath spins S in the absence of donor electrons,  assumes the form

α 
 = g ∗ μB B + (M 0 + bq ηq ). (10)
g0 μB q

Here g ∗ is the electron effective Landé factor; bq and ηq are Fourier components
of |ψ(r)|2 and of bath magnetization fluctuations M(r) − M 0 , respectively. If the
electron localization length is much longer than an average distance between the
bath spins, the summation over q can be extended to infinity, but otherwise an
appropriate cutoff qmax should be implemented.
Except in the immediate vicinity of the spin ordering temperature, the probability
distribution of ηq is, to a good approximation, Gaussian for any mixed state that
can be described by spin temperature with variance, according to the fluctuation-
dissipation theorem, determined by an appropriate integral over ω of the imaginary
part of χq (ω). Since  is linear in ηq , the central limit theorem implies that the
probability distribution of  in the absence of the electron, PS (), is also Gaussian:

PS () = ZS−1 exp[−(


z −
0 )2 /2σ 2 − 2⊥ /2σ⊥2 ], (11)

where ZS = (2π )3/2 σ σ⊥2 is the probability-normalizing constant insuring that



dPS () = 1 and
2 =
2z + 2⊥ with the z-axis taken along the magnetic
field. The three parameters characterizing cubic systems are then given by

α

0 = g ∗ μB B + Mo , (12)
g0 μB
 2 
α
2
σ (⊥) = kB T χ̄ (⊥)q |bq |2 /μ0 . (13)
g0 μB q
944 A. Bonanni et al.

If energies of bath magnetic excitations are smaller than the thermal energy kB T
and their q-dependencies are irrelevant, then χ̄ (⊥)q (T , H ) are given in Eq. 7. Far
from magnetization saturation, χ⊥ = χ = χ .
By adding the free energy of the electron, so that

P() = Z −1 2 cosh(
/2kB T )PS (), (14)

the BMP effects in thermal equilibrium are taken into account with the partition
function Z being determined by the probability normalization condition. Accord-
ingly, the BMP free energy becomes F = −kB T ln(Z/ZS ), from which the
BMP contribution to magnetization and specific heat is obtained. Furthermore, by
minimizing F in respect to the BMP localization radius, self-trapping effects can
be assessed [95]. Integrating P() over  directions, one obtains the probability
distribution of the electron spin-splitting
that can be directly probed by a
variety of optical and magnetic resonance experiments. In particular, theoretically
determined P(
) describes quantitatively the shape, width, and position of the
spin-flip Raman scattering line in a number of n-type DMSs [96]. Additional
inhomogeneous spin-splitting broadening caused by statistical fluctuations of com-
position x (chemical disorder), leading to different
0 values in particular BMPs,
can also be incorporated into the theory [42].
In a paramagnetic phase and in the absence of an external magnetic field, P(
)
attains a maximum at
= 2εp and (8kB T εp )1/2 for εp kB T and εp kB T ,
respectively, where the polaron energy

α 2 χ (T )
εp (T ) = , (15)
32πg02 μ2B aB3

where aB is an effective Bohr radius of the relevant carrier localized state. As seen,
spontaneous spin-splitting appears even in the paramagnetic phase. It is dominated
by the polaron effect, i.e., by bath spin polarization produced by the carrier spin
at low temperatures, whereas thermodynamic magnetization fluctuations lead to
a molecular field whose contribution to
takes over in the high-temperature
regime.
Since non-scalar spin-spin interactions, such as Dzyaloshinskii-Moriya coupling,
break spin momentum conservation, the BMP formation time is of the order of T2(S)
of the bath spins. Hence, up to a time scale of T1(S) , adiabatic rather than isothermal
magnetic susceptibilities describe energetics of the system [97]. Furthermore, by
solving the quantum Liouville equation with the Hamiltonian Hs = ŝ ·  and
averaging the resulting spin-density matrix with P(), one obtains information on
dephasing of the electron spin for t < T2 at arbitrary values of M0 (T , H ) and
magnetic field B [47].
18 Dilute Magnetic Materials 945

Quantum Localization and Mesoscopic Phenomena: Colossal


Magnetoresistance

In addition to contributing to the momentum relaxation rate, spin-disorder scattering


plays an essential role in controlling the magnitude of quantum coherence effects in
charge transport ( Chap. 5, “Quantum Magnetism”), particularly in metals, where
the DOS and, thus, 1/τsi are relatively large. Even residual magnetic impurities
or spin-carrying defects in nominally ultra-pure noble metals perturb quantum
interference phenomena below 100 mK [98].
The presence of giant spin-splitting affects in a dramatic way quantum localiza-
tion [31] and mesoscopic phenomena [99] in DMSs. Also spin-disorder scattering
in the paramagnetic phase, presumably involving preformed FM clusters, enhances
localization and leads to colossal negative magnetoresistance (CMR) in the vicinity
of TC [38, 31] ( Chap. 9, “Magnetotransport”). The presence of CMR is a
trademark of magnetic and dilute magnetic semiconductors. It allowed testing of
dynamic renormalization group equations at the magnetic field-induced Anderson-
Mott insulator-to-metal transition [100]. At the same time, quantitative description
of the mutual relationship between magnetism and Anderson-Mott localization is
one of open issues in the physics of dilute magnetic materials.

Interplay of sp–d(f ) Exchange Interactions and Spin-Orbit Coupling

This interplay is a core resource for spintronic applications of magnetic materials,


accounting for a range of magnetotransport [AMR, AHE ( [101],  Chap. 9, “Mag-
netotransport,” 6)] and magnetooptic ( Chap. 10, “Magneto-optics and Laser-In-
duced Dynamics of Metallic Thin Films”) phenomena as well as the magnitude of
critical currents for spin-transfer [102], spin-orbit torque [103], and spin-locking in
topological materials [104,82]. In the spin-orbit Hamiltonian Hso = μB s·(E×v)/c,
the relevant electric field E originates from atomic and impurity potentials, lack of
inversion symmetry in some crystal structures or at interfaces, space-charge layers,
or external sources, whereas the required velocity v from drift motion, pseudo-
momentum k, atomic orbital motion, and a real velocity in a crystal, often described
in terms of the Berry curvature dependent on the topology of a given Brillouin zone
( Chap. 5, “Quantum Magnetism”).
A number of these functionalities, particularly those insensitive to disorder,
can be examined numerically by exploiting progress in the development of fully
relativistic codes ( Chap. 10, “Magnetooptics and Laser-Induced Dynamics of/in
Thin Metalic Films”). In the case of DMSs, particularly rewarding and numerically
efficient has been the application of semiempirical multiband kp and multiorbital
tight-binding approaches taking into account the presence of magnetic impurities
within VCA and MFA [23, 24], the methodology generalized more recently
946 A. Bonanni et al.

to predict the anomalous quantum Hall effect in topological materials [84, 51]
(sections “p–d Zener Model” and “Dilute Ferromagnetic Insulators and Topological
Insulators”). The incorporation of spin-orbit effects into theories of spin-disorder
scattering, magnetic polarons, strong coupling, and quantum localization, particu-
larly in the case of overlapping p-type subbands, is technically more demanding.

Dominant Spin-Spin Interactions

Dipole-Dipole Interactions

Dipolar interactions are ubiquitous in nature and due to their long-range may domi-
nate other types of interactions in the limit of low magnetic impurity concentrations,
particularly in magnetic insulators with rare earth impurities, where competing
exchange coupling decays rapidly with spin-spin distance. Due to randomness, the
resulting ground state is a spin glass, as discussed in section “Spin-Glass Systems”.

Direct Spin-Spin Interactions

We consider pairs of interacting spins. In contrast to the direct Coulomb interaction


(the Hartree term) that decays as inverse distance, the exchange coupling between
spins (the Fock term) is nonzero (at least neglecting spin-orbit effects) only if
there is an overlap between the relevant wave functions. This is the case of
molecular dimers, such as Mn2 , for which the AF exchange integral comes from
the kinetic exchange (quantum hopping between d orbitals of the two magnetic
atoms, i.e., the direct interaction). The determined value of J for the molecular
dimer Mn2 [105] is shown in Fig. 4 in comparison to AF exchange energies
JNN of the nearest-neighbor (NN) Mn, Co, and Eu cation dimers in II-VI and
IV-VI families of DMSs. The data for these impurities in various hosts point to a
relatively slow decay of AF JN N with distance rN N . Furthermore, −JN N at given
rNN is substantially larger for tetrahedral environment (II-VI DMSs) compared to
the octahedral case (IV-VI DMSs). This suggests the presence of indirect, i.e.,
anion-mediated, exchange mechanisms in the case of TM impurities in compound
semiconductors.
In agreement with the expectation, the smallest magnitudes of AF exchange
energies are observed in the case of Eu cation dimers, for which spins reside on the
highly localized 4f shells. Surprisingly, however, in the end compounds, i.e., in Eu
monochalcogenides, the nearest-neighbor interaction is FM. According to Kasuya’s
model, in the presence of FM spin arrangement, d–f splitting of the conduction
band lowers the energy denominator for quantum hopping between f and d states
on the NN Eu ions, JNN ≈ xJdf |Vdf |2 /(Ed − Ef )2 . In the case of f orbitals and
for x = 1 (i.e., giant splitting of the conduction band), this interaction is expected
to be stronger than the competing x-independent AF coupling.
18 Dilute Magnetic Materials 947

100

ZnS
ZnTe ZnSe
Exchange energy - JNN (K)

CdSe
10 CdTe
CdS Mn2 ZnO
HgSe
HgTe

PbSe
triangles: Co
1
squares: Mn
PbTe circles: Eu
PbS

0.1
10 20 30 40
Cation concentration N0 (nm )
-3

Fig. 4 Compilation of experimentally determined antiferromagnetic exchange integrals −JNN for


the nearest-neighbor Mn, Co, and Eu cation dimers in various DMSs as a function of the cation
concentration N0 , i.e., the inverse cation dimer distance for fcc lattices rNN = (4/N0 )1/3 /21/2 ,
as collected by Shapira [106], Savoyant et al. [107], and Story [108]. In the case of the wurtzite
compounds, values corresponding to two nonequivalent dimers are shown. The convention, Hij =
−2Jij S i ·S j , is employed. The magnitude of −J for Mn2 molecular dimer is also presented [105]

Superexchange

This indirect coupling between localized spins exists in virtually all compounds
with localized spins and typically dominates in nonmetallic materials containing
transition metals (TMs). The superexchange ( Chap, 2, “Magnetic Exchange
Interactions”) proceeds via hybridization of magnetic levels with orbitals of atoms
located between the spins in question, usually, but not always, p orbitals of relevant
anions. For this indirect interaction, within the lowest-order perturbation theory
in the hybridization, the interaction Hamiltonian is bilinear, Hij = −2S i Jˆij S j .
Components of the exchange tensor Jˆij are proportional to |Vkd(f ) |4 and decay fast
with the distance rij corresponding to NN, NNN, NNNN, ... spin pairs, typically
on average exponentially, though not necessarily monotonically, as the direction of
r ij in respect to the orientation of molecular orbitals is relevant, too. The scalar
(isotropic) part of Jˆij [i.e., Jij I , where I is the identity (unit) matrix] usually
dominates though the remaining components brought about by spin-orbit coupling,
particularly the Dzyaloshinskii-Moriya antisymmetric term, Hij DM = −2D DM ·(S ×
i
S j ), play often an important role in media with inversion asymmetry, also associated
948 A. Bonanni et al.

with proximity to an interface. In general, group theory determines the form, if not
the magnitude, of Jˆij for particular spin pairs and hosts.

DMSs with Transition Metals


Antiferromagnetic superexchange: A vast majority of nonmetallic TM compounds
are antiferromagnets or ferrimagnets. Figure 4 presents the nearest-neighbor (NN)
AF exchange energies JNN for Mn and Co dimers in II-VI and IV-VI DMSs,
as determined from inelastic neutron scattering and magnetization steps in high
magnetic fields in samples with a few percent of TM ions, as well as by modeling
dependencies of magnetization and specific heat on a magnetic field and tempera-
ture. These experiments, particularly magnetization steps [106], provided also some
information on Jij for more distant neighbors as well as, together with the EPR
linewidth studies, on the Dzyaloshinskii-Moriya term whose magnitude increases
on going from sulphides to selenides and tellurides [109]. In general, the employed
ab initio and tight-binding approaches [73, 110, 107] describe satisfactorily the
observed sign and magnitude of JNN .
Ferromagnetic superexchange: According to the Goodenough-Kanamori-
Anderson rules ( Chap, 2, “Magnetic Exchange Interactions”), superexchange is
FM for certain charge states of TM ions and bond arrangements. It was found
employing a tight-binding approximation that Jij > 0 in the case of Cr2+
[111] and Mn3+ [33] ions in a tetrahedral environment. These expectations are
corroborated by experimental results for (Ga,Mn)N, as discussed in section “Dilute
Ferromagnetic Insulators and Topological Insulators” (Fig. 11). According to ab
initio studies of Bi and Sb tetradymite chalcogenides [56], the coupling is FM
for pairs of Ti, V, Cr, and Mn spins, whereas it assumes an AF character for
Fe and Co pairs. If the computational approach properly describes the localized
nature of magnetism in these systems, the determined chemical trend indicates that
superexchange is the relevant interaction mechanism.

IV-VI DMSs with Rare-Earth Metals


In the most-studied Eu monochalcogenides, where spin S = 7/2 resides on the
highly localized 4f shell, magnetic properties can be described in terms of FM
NN and AF next NN (NNN) exchange integrals, J1 and J2 , respectively. Pressure
studies revealed a strong dependence of Ji values on the lattice parameter and,
−20 −10
thus, on the Eu pair distance rij , J1 (rNN ) ∝ RNN and J2 (rNNN ) ∝ rNNN [112].
As mentioned in section “Direct Spin-Spin Interactions”, according to Kasuya’s
model, the direct Eu-Eu interaction involving quantum hopping between f and d
states is relevant. At larger distances indirect AF coupling via anions takes over,
which determines the magnitude of J2 , in agreement with the presence of p–f
hybridization revealed by ab initio computations [113]. As shown in Fig. 4, in Eu-
doped lead chalcogenides, only antiferromagnetic coupling is observed [108], as the
magnitude of Kasuya’s mechanism diminishes with decreasing x (section “Direct
Spin-Spin Interactions”). A resonant enhancement of AF f –f coupling occurs
when the Fermi level is pinned by donor 5d states degenerate with the valence band
in p+ -(Sn,Gd)Te [70].
18 Dilute Magnetic Materials 949

Carrier-Mediated Spin-Spin Coupling: Intra- and Interband


Contributions

RKKY Interaction
As shown by Ruderman, Kittel, Kasuya, and Yosida, under the presence of a sp–
d(f ) exchange interaction, a localized spin S i generates a spin polarization of the
carrier sea that affects another localized spin S j ( Chap, 2, “Magnetic Exchange
Interactions”). Since spin polarization shows Friedel oscillations (associated with
the Kohn anomaly due to the Fermi energy cutoff), the sign of resulting bilinear
coupling energy Jij , as given by the lowest-order perturbation theory in Jsp−d(f ) ,
changes between FM and AF as a function of rij with period λF /2, where λF is the
de Broglie wavelength of carriers at εF ; the period is short in metals but substantially
longer than an average distance between spins in doped semiconductors, where
carrier density typically n < xN0 . The interaction amplitude is proportional to
(Jsp−d(f ) /N0 )2 and the density of carrier states at εF and decays as rij−d , where d
is the DOS dimensionality. Similarly to superexchange, spin-orbit interactions give
rise to the presence of non-scalar terms in spin-spin coupling.

Bloembergen-Rowland Mechanism
This mechanism takes into account the presence of spin polarization of occupied
bands generated by localized spins. Since here the perturbation theory involves
interband matrix elements, this coupling is expected to be particularly strong in the
inverted band structure cases of topological materials, such as (Hg,Mn)Te [114] and
(Bi,Cr)2 Se3 [51]. In the case of (Hg,Mn)Te and related alloys and to the lowest order
in perturbation theory that leads to bilinear coupling, Jij is ferromagnetic for the NN
γ
and antiferromagnetic for more distant pairs for which it decays as 1/rij , where γ
varies between 4 and 5 in the HgTe type of systems [114], a much slower decrease
with rij compared to superexchange. Furthermore, this mechanism is relevant for p-
type tetrahedrally coordinated DMSs, in which quantum hopping between various
valence subbands is allowed [55].

p–d Zener Model


In the frame of the superexchange, RKKY, and Bloembergen-Rowland approaches,
the thermodynamic properties of the system are inferred from the eigenvalues of the
Hamiltonian which contains the relevant bilinear interactions S i Jˆij S j for all pairs
of localized spins at a given value of the Fermi energy of the spin-unpolarized host.
According to the p–d Zener model, implemented with mean-field and molecular-
field approximations, one looks for the spin magnetization of localized magnetic
ions M(r) that minimizes – for given T , H , and carrier density p – the total
free energy functional F[M(r)]. This functional includes (i) a term related to the
localized spin in the absence of carriers FS [M(r)] and (ii) the carrier contribution
Fc [M(r)] in the presence of magnetization produced by localized spins [23].
If the energetics is dominated by spatially uniform magnetization M and the
magnetic ions carry no orbital momentum (L = 0), the free energy density of
localized spins in the magnetic field H is expressed as
950 A. Bonanni et al.

 M
FS [M](T , H ) = dM o · h(M o ) − M · H . (16)
0

Here, h(M o ) is the inverse function of M o (h), where M o is the experimental


macroscopic magnetization of the localized spins in the absence of carriers in the
magnetic field h and at temperature T . In this way, the contribution due to intrinsic
spin-spin interactions, like superexchange, is taken into account.
The carrier-related term Fc [M(r)] is assessed within the multiband kp [79] or
multiorbital tight-binding theory applicable for a given host including the sp–d(f )
coupling and the spin-orbit effects. The incorporation of spin-orbit interactions is
imperative [19], as – together with single-ion anisotropy – it accounts for the mag-
netic anisotropy and for other specific properties of the magnetic systems. Within
this formulation, intra-band and interband contributions (involving the RKKY and
Bloembergen-Rowland mechanisms, respectively) are inherently accounted for. The
model is valid for arbitrary M and allows taking account of strain, confinement, the
presence of boundary states in topological materials, and the Landau quantization
of the carrier spectrum. The carrier correlation effects are tied in, by introducing
a Fermi-liquid-like parameter AF , which dilates the Pauli spin susceptibility of the
hole liquid.
Such effects would be arduous to treat within the RKKY and the Bloembergen-
Rowland approaches for three reasons. First, the spin-orbit coupling leads to
non-scalar terms in the spin-spin Hamiltonian. Second, the host is assumed to be
spin unpolarized, the approximation breaking down in a ferromagnetic phase. Third,
there is an increasing amount of evidence that a description of magnetism in terms of
a bilinear exchange Hamiltonian is inadequate if the range of spin-spin interactions
is greater than the average distance between spins, as in the case of DFSs.
The implementation of the p–d Zener model for the determination of various
DFS properties and functionalities is now well established, at least for collinear
magnetism and for systems not too close to the metal-insulator transition [23, 24].
Here below, we summarize the p–d Zener model of TC [19]; it is extensively
reviewed elsewhere [23].

Theory of Curie Temperature


Near the Curie temperature TC and at H = 0, where M is small and the free energy
is an even function of M, one expects Fc [M] − Fc [0] ∼ −M 2 . It is convenient
to parameterize this dependence by a generalized carrier spin susceptibility χ̃c ,
which is related to the magnetic susceptibility of the carrier liquid according to
χc = AF (g ∗ μB )2 χ̃c . In terms of χ̃c ,

Fc [M] = Fc [0] − AF χ˜c β 2 M 2 /2(gμB )2 . (17)

By expanding BS (M) for small M and introducing the spin susceptibility of the
magnetic ions in the absence of carriers, χ̃S = χ /(gμB )2 , one arrives at the mean-
field formula allowing to determine TC :
18 Dilute Magnetic Materials 951

AF β 2 χ̃S (TC , q)χ̃c (TC , q) = 1, (18)

where β should be replaced by the s–d exchange integral α in the case of electrons
(section “Exchange Interactions Between Band and Localized Spins”) and q denotes
the Fourier component of the magnetization texture, for which TC attains the
highest value. It is convenient to parametrize experimental values of χ̃S by χ̃S =
xeff N0 /S(S + 1)/(T − Θp ), where xeff = x and Θp = 0 take into account coexisting
superexchange interactions. For the spatially uniform magnetization (q → 0), one
then obtains

TC = TF + Θp , (19)

where TF is given by

TF = xeff N0 S(S + 1)AF χ̃c (TC )β 2 /3kB . (20)

The magnitude of χ̃c appearing in Eq. 18 can be determined from the linear
response theory. Assuming that Jsp−d /N0 ≡ β in Eq. 4, so that only anion p bands
are relevant [55]:

 |ui,k |sM |uj,k+q |2 fi (k)[1 − fj (k + q)]


χ̃c = lim 2 , (21)
q→0 Ej (k + q) − Ei (k)
ij k

where the term i = j describes the intraband (RKKY-like) coupling, whereas the
interband (Bloembergen-Rowland-like) interactions correspond to i = j terms; sM
is the component of the spin operator along the direction of magnetization; ui,k is the
periodic part of the Bloch wave function; and fi (k) is the Fermi-Dirac distribution
function for the i-th band.
In thin films, heterostructures, and superlattices, owing to the formation of inter-
facial space charge layers, the hole density and corresponding Curie temperatures
TC [p(z)] are nonuniform even for a uniform distribution of acceptors and donors.
The role of nonuniformity in the carrier distribution grows on reducing the thickness
t of magnetic layers and is particularly relevant in those structures where p(z) can
be tuned electrostatically, for instance, by the gate voltage. When t is larger than the
phase coherence length Lφ , the region with the highest TC value determines TC of
the whole structure. If, however, t < Lφ , the value of local magnetization M(z) and
TC are determined by the distribution p(z) across the whole channel thickness. In
this regime two situations are relevant.
If disorder is strong  < t, the scattering broadening makes dimensional
quantization irrelevant though quantum mechanical non-locality remains important.
Under these conditions the magnitude of a layer’s TF can be expressed as [115]
  
TF = dzTF [p(z)] dzp(z)2 /[ dzp(z)]2 , (22)
952 A. Bonanni et al.

where TF (p) is to be determined from the relevant 3D model and p(z) is to be


evaluated from the Poisson equation taking into account the pinning of the Fermi
energy by surface states.
The opposite limit of weak disorder,  t, was also considered [116,22]. Owing
to a typically large confinement-induced splitting between heavy- and light-hole
subbands, only one ground-state heavy-hole subband is occupied, for which the
p–d exchange is of the Ising type, Hpd = −N0 βsz Sz , so that

TF = N0 xeff S(S + 1)AF β 2 m∗ /(12π h̄2 kB L̃W ), (23)

Here m∗ denotes the in-plane effective mass, and L̃W is an effective width of the
region occupied by carriers and spins relevant to ferromagnetism given in Eq. 8 [22,
117]. As seen, owing to a steplike form of DOS in the 2D case, TF does not depend
on the hole density in this case. The expression for TF was generalized further to the
case of arbitrary degeneracy of the hole liquid and by including effects of disorder
via scattering broadening of DOS [118].
Theoretical approaches were developed to allow evaluation of the Curie temper-
ature for ferromagnetic ordering of magnetic impurities mediated by Dirac carriers
at the surface of 3D topological insulators [119, 120]. The Ising type of exchange
was assumed, Hex = −N0 Jz sz Sz , leading to a gapped dispersion given by

ε(k) = ±[(Jz M/2gμB )2 + (h̄vf k)2 ]1/2 , (24)

where vf is the Fermi velocity. For such a case [119], in our notation,

TF = N0 xeff S(S + 1)AF rJz2 (Ec − |εF |)/(24π h̄2 kB vf2 L̃W ), (25)

where r is the number of Dirac cones at a given surface, Ec is a cutoff energy


associated with the termination of the Dirac surface band, and L̃W is the penetration
depth of Dirac carriers related to their envelop function according to Eq. 8.
The formation of spin-density waves is expected in the case of carrier-mediated
ferromagnetism in 1D systems [117].

Magnetic Properties of Dilute Magnetic Materials

Spin-Glass Systems

The coexistence of randomness and frustration in spin-spin interactions ( Chap, 2,


“Magnetic Exchange Interactions”) results in a distinct magnetic behavior known as
spin-glass freezing. The randomness is typically brought about by dilution (frozen
disorder), and therefore the spin glass is a low-temperature magnetic phase of a
number of dilute magnetic materials [10].
18 Dilute Magnetic Materials 953

A well-known example of a frustrated system is a triad of antiferromagnetically


interacting spins located at corners of a triangle. Accordingly, spin-glass freezing
is observed in, for instance, zinc-blende or wurtzite II-VI Mn-based DMSs such
as Cd1−x Mnx Te, x  0.65 in which AF superexchange (section “DMSs with
Transition Metals”) is the dominant exchange mechanism for all Mn-Mn distances,
and the end compounds (x = 1) are AF [54, 106]. To the spin-glass family belong
also other types of DMSs such as rock salt Sr1−x Eux S with x  0.5 [121] and
spinel Cd(In1−x Crx )2 S4 with x  0.85 [122], in which the frustration is enforced
by a competition between different interactions: the NN and the NNN exchange
coupling is FM and AF, respectively, and the end compounds (x = 1) are FM
(section “IV-VI DMSs with Rare-Earth Metals”). Another example is tetragonal
LiHox Y1−x F4 with x  0.20 [123], characterized by (i) strong single-ion magnetic
anisotropy that turns Ho ions into Ising spins, (ii) weak exchange coupling between
them meaning that magnetism is driven by dipolar interactions that are FM or AM
depending on the position of a given Ho pair in respect to the c-axis, and (iii) FM
ordering at x = 1. Actually, spin-glass freezing was discovered and most thoroughly
studied, not in semiconductors, but in dilute magnetic metals [10], such as CuMn,
in which the RKKY exchange interaction operates (section “RKKY Interaction”).
Despite their diversity, these materials exhibit a similar set of spin-related
properties [10, 122, 12], including:

(1) A characteristic cusp in the temperature dependence of magnetic susceptibility


χ (T ), whose maximum in the static and H → 0 limits defines the freezing
temperature Tf (Fig. 5);
(2) Divergence of a coefficient of nonlinear magnetic susceptibility anl =
limH →0 ∂ 2 χ (H )/∂H 2 at T → Tf+ ;
(3) Broadening and shift of the cusp position to higher temperatures with the
magnitude and frequency f of the magnetic field H , respectively;
(4) A thermal equilibrium behavior of field-cooled magnetization MFC (T , H ) that
shows no thermal hysteresis when quenching below Tf and annealing above Tf
in a constant magnetic field H ;
(5) Out-of-equilibrium magnetization dynamics below Tf in response to field
changes, which fulfils the relation MFC (T , H = 0, t) + MZFC (T , H = 0, t) =
MFC (T , H ), where MFC (T , H = 0, t) and MZFC (T , H = 0, t) denote time-
dependent magnetizations after removing and applying the magnetic field H ,
respectively, and the zero-field-cooled magnetization MZFC (T , 0, t) ≡ 0 in a
spin-glass case (Fig. 5);
(6) Other slow-dynamic glassy phenomena at T < Tf including: (i) aging, a
dependence of the magnetization evolution rate on the waiting time, when
the system was kept at given T < Tf and H , before H was changed; (ii)
rejuvenation, by partial erasing of aging by abrupt lowering of temperature;
and (iii) memory, the partial erasing of rejuvenation by returning to initial
temperature;
(7) No changes in the spin structure factor S(q, T ) or resistivity ρ(T ) on cross-
ing Tf ;
954 A. Bonanni et al.

Fig. 5 Magnetization in the vicinity of spin-glass freezing temperature Tf in CdCr1.7 In0.3 S4 .


Zero-field-cooled (ZFC), field-cooled (FC), and thermoremanent (TMR) magnetizations in 10 Oe
(10 kA/m/4π ) are shown. (From Vincent et al. [122])

(8) No anomaly in the temperature dependence of spin-specific heat at Tf but rather


a broad maximum at T > Tf .

According to the present understanding [12] and in contrast to dielectric window


glasses, spin glasses undergo at Tf a continuous phase transformation to a distinct
thermal equilibrium spin state. This new phase is characterized by a nonzero value
of the Edwards-Anderson order parameter:

qEA = lim S i · S i (t), (26)


t→∞

where the bar denotes the configuration average with respect to random spin
positions and the brackets denote the thermal or dynamical average. The existence
of a phase transition at nonzero temperatures is supported by a careful experimental
examination of static [124] and dynamic [125, 45] critical scaling behavior of
magnetization as function of T , H , and f in the immediate vicinity of Tf for AgMn,
Fe0.5 Mn0 .5 TiO3 , and (Cd,Mn)Te, respectively.
Another relevant question is whether the lower critical dimensionality for
Heisenberg spins is dl < 3, so that the observed phase transition in 3D cases is
driven by magnetic anisotropy that effectively makes the spins Ising-like. Figure 6
shows Tf of CuMnPt and AgMnAu as a function of a magnetic anisotropy parameter
A that increases with the concentration of Pt and Au impurities bringing a sizable
spin-orbit interaction to this Heisenberg spin system [126]. As seen, the data point
to nonzero values of Tf in the limit A → 0. A similar conclusion emerges from
Fig. 7 showing comparable magnitudes of Tf in Mn-doped II-VI selenides and
tellurides, despite the strength of the Dzyaloshinskii-Moriya interaction differing
18 Dilute Magnetic Materials 955

35

Freezing temperature Tf (K)


CuMnPt
30 AgMnAu

25

20

15
0.0 0.1 0.2 0.3 0.4
Magnetic anisotropy parameter A0.8

Fig. 6 Spin-glass freezing temperature Tf in CuMn4% and AgMn5.5% as a function of a magnetic


anisotropy parameter A to power 0.8, changed by co-doping with Pt (concentration between 0 and
1.96%) and with Au (concentration between 0% and 5%), respectively. Straight lines are guides
for the eye showing that Tf > 0 for A → 0. (Adapted from Fert et al. [126])

by a factor of about two for these two families of Heisenberg DMSs [127].
Altogether these findings suggest that the spin-glass phase transition occurs also
for the Heisenberg case. However, somewhat surprisingly critical exponents appear
similar for Heisenberg [45] and Ising [125] spins.
A great deal of information on spin glasses comes from Monte Carlo numerical
simulations [11, 12]. Within the Edwards-Anderson model, classical spins residing
in a cubic lattice are coupled by NN exchange interactions Jij , whose magnitudes
are randomly selected from a Gaussian distribution with a zero mean value. There is
a converging view that for such a model Tf > 0 for both Ising and Heisenberg spins
in the 3D case [12, 128]. There is also a growing amount of evidence that spin-glass
freezing of Heisenberg spins is accompanied by a chiral order, i.e., by a nonzero
value of the chiral order parameter:

(μ)
qchiral = S i−μ̂ · (S i × S i+μ̂ ), (27)

where μ̂ is the unit lattice vector along x, y, or z [12, 128]. The presence of chiral
ordering is thought to account for the lack of proportionality between the anomalous
Hall resistivity and magnetization observed in spin glasses, such as AuMn [129].
Since a number of optimization problems can be mapped onto the determination
(z)
of thermal equilibrium spin orientations {Si } for a given set of interaction
magnitudes {Jij }, progress in spin-glass computation tools has had a significant
impact on other fields [13]. A particularly timely question concerns the effectiveness
956 A. Bonanni et al.

x
1.0 0.064 0.0156

Zn1-xMnxSe
Zn1-xMnxTe
Cd1-xMnxTe
10 Cd1-xMnxSe
Hg1-xMnxTe
Tf(K)

Hg1-xMnxSe
(Cd1-xMnx)3As2
1

0.1
1.0 1.5 2.0 2.5 3.0 3.5 4.0

x-1/3
Fig. 7 Spin-glass freezing temperature Tf in various bulk II1−x Mnx -VI DMSs and
(Cd1−x Mnx )3 As2 as a function of x −1/3 (points), as implied by the percolation theory for
spin-spin couplings decaying exponentially with the pair distance [34]. Straight line is a guide for
the eye. Relatively high values of Tf for inverted band structure semiconductors Hg1−x Mnx Te(Se)
at x  0.1 and (Cd1−x Mnx )3 As2 point to the presence of an additional exchange interaction,
presumably ferromagnetic interband Bloembergen-Rowland (Van Vleck) contribution (sections
“Bloembergen-Rowland Mechanism” and “p–d Zener Model”). (The data points collected by
Gała̧zka et al. [54])

(z)
of quantum annealing procedures [14] that aims at accelerating the search for Si

by adding a transverse field Hx , which means that i S (z) zi ceases to be a constant
of motion and, thus, the ground state {Si(z) } for Hx → 0 can be achieved by quantum
tunneling.
On the theoretical side, two approaches have generated a considerable interest.
One is the Sherrington-Kirkpatrick model, in which the probability distribution
P(Jij ) is Gaussian but the range of interactions between classical Ising spins is
infinite, so that a mean-field approach should be valid. Its implementation proceeds
by employing a replica trick allowing, in the configurational averaging of the free
energy over P(Jij ), the replacement of ln Z[Jij Si(z) Sj(z) ] by (Z[Jij Si(z) Sj(z) ])n ,
i.e., by n replicas of the system. In the time-honored and exact solution proposed
by Parisi, order parameters involving different pairs of replicas (actually their
overlap) are not identical (replica symmetry breaking – RSB). Within the RSB
model, wandering of the coupled spin system between closely lying local free
energy minima accounts for glass-like dynamics (hierarchical dynamics) [15]. An
18 Dilute Magnetic Materials 957

Almeida-Thouless line separates the paramagnetic and spin-glass phase in the H –T


plane.
Another proposal, the droplet theory [16], is developed for short-range interac-
tions. In this picture, there exists a unique spin ground state (and its time-reversal
counterpart), whereas local deviations from it, in the form of compact and spatially
uncorrelated droplet-like regions, constitute low-energy excitations accounting for
the spin-glass dynamics. Exchange stiffness of spins adjacent to the droplet surface
determines the droplet energy E that, because of randomness and frustration,
increases rather slowly with the droplet radius R, E ∼ R θ , where θ < d/2.
As the droplet magnetic moment m ∼ R d/2 , an arbitrarily small magnetic field
H magnetizes droplets with appropriately large radii R ∼ H −2/(d−2θ) and, thus,
destabilizes the spin-glass state.
One of the signatures of glassy dynamics is the 1/f noise characterized by
a power spectrum S(f ) ∼ 1/f γ , where 1  γ  2 is found experimentally
in a variety of systems and expected theoretically within a number of disordered
media models. While magnetization noise was detected directly in spin glasses
[130], particularly accurate information was obtained by exploiting an outstanding
sensitivity of mesoscopic electronic systems to spin configurations [131, 46]. As
expected, studies of resistance as a function of time reveal 1/f noise over a wide
frequency range in nanostructures of CuMn [131] and (Cd,Mn)Te [46] below Tf .
However, more insight into the nature of spin glasses is provided by examining
the Fourier transform of the time dependence of S(f ) determined over several
consecutive time intervals (say a series of 70 S(f )’s, each extracted from resistance
noise collected over consecutive 1-h intervals). Such a second spectrum s (2) (f ) is
white for independent fluctuators (Gaussian noise) but assumes a nonwhite character
if events of spin reconfigurations are correlated. As shown in Fig. 8, the fluctuators
appear independent for Cd0.93 Mn0.07 Te [46]. However, some correlations are found
for Cd0.8 Mn0.2 Te [46], and they are significantly stronger for Cu0.91 Mn0.09 [131].
These results lead to the conclusion that the hierarchical model of the spin-glass
phase applies to dilute magnetic metals, in which spin-spin interactions are long-
range, whereas the droplet picture is more appropriate for DMSs with short-range
antiferromagnetic coupling.

p-Type Dilute Ferromagnetic Semiconductors

In virtually all semiconductor families, the coexistence of hole carriers with dilute
Mn2+ ions results in ferromagnetic ordering at sufficiently low temperatures.
This ferromagnetism is mediated by itinerant or weakly localized holes that are
strongly coupled to Mn spins via p–d hybridization. In DFSs of III-V [(Ga,Mn)As,
(In,Mn)As [20]] and V-VI [(Bi,Mn)2 Te3 [133]] compounds and also of elemental
semiconductors [(Ge,Mn) [134]], the holes are provided by Mn ions themselves,
which act as acceptors. Deviation from stoichiometry, such as cation vacancies, is
another source of holes, the case of IV-VI DFSs [(Ge,Mn)Te [28,29], (Pb,Sn,Mn)Te
958 A. Bonanni et al.

1.5

1.0

0.5
Log second spectral density (Hz-1)

Cd0.8 Mn0.2Te

0.0 Cd0.93 Mn0.07Te


Gaussian background

-0.5
10-3 10-2 10-1

2.0
Cu0.9 Mn0.1
Cd0.8 Mn0.2 Te
1.5

1.0

0.5

0.0
10-3 10-2 10-1
Frequency f (Hz)
(2,φ)
Fig. 8 Upper panel: Second noise spectra in the phase domain sf of (Cd,Mn)Te quantum wires
at 50 mK in the bandwidth from 0.1 to 0.6 Hz [46] (points) are compared to expectations for
(2)
the Gaussian 1/f noise (solid line). Lower panel: Frequency dependence of sf (normalizing to
and subtracting the Gaussian expectation) for Cd0.8 Mn0.2 Te at 50 mK [46] and Cu0.91 Mn0.09 at
11 K [132] (points). Lines, except for the Gaussian background, are guides for the eye. (From
Jaroszyński et al. [46])

[21]] and also of I-II-V systems [Li(Zn,Mn)As [135]]. In certain compounds holes
can be introduced by co-doping with anion- or cation-substitutional acceptors [e.g.,
(Cd,Mn)Te/(Cd,Mg,Zn)Te:N [22] and (K,Ba)(Zn,Mn)2 As2 [30], respectively]. In
the case of (Ge,Mn)Te, ferromagnetism coexists with ferroelectricity [136].
These systems share a number of common features affecting their magnetic
properties.
First, the hole-mediated interaction competes with a short-range antiferromag-
netic superexchange, so that there is an optimum Mn content for achieving the
highest TC (so far reaching ∼200 K in (Ga,Mn)As, (Ge,Mn)Te [28, 29], and
(K,Ba)(Zn,Mn)2 As2 [30]) which according to magnetization saturation corresponds
to xeff  0.1.
18 Dilute Magnetic Materials 959

Second, these materials show a strong tendency to phase separation (see sec-
tion “Heterogenous Magnetic Semiconductors and Oxides”) – dilute Mn spins can
coexist with nanocrystals of, for instance, ferromagnetic Mn-rich (Mn,Ga)As in
(Ga,Mn)As [137] or antiferromagnetic MnTe in (Ga,Mn)Te [29].
Third, in addition to the solubility limit of Mn ions, there exists a self-
compensation mechanism with generation of donor defects (e.g., Mn interstitials
or As antisites) that limit the achievable hole density (section “Heterogenous Mag-
netic Semiconductors and Oxides”). These and other defects may alter magnetic
properties.
Fourth, as in any doped semiconductor, carriers in DFSs undergo the Anderson-
Mott metal-to-insulator transition (MIT) at pc ≈ (0.26/aB )3 , where the effective
Bohr radius aB is diminished by coupling to localized spins, rather substantially
in the strong coupling limit (section “Spin-Splitting of Extended States: Strong
Coupling”). Nanoscale critical fluctuations in the carrier density on approaching
and crossing the MIT result in the appearance of spin-glass (section “Spin-Glass
Systems”) and superparamagnetic regions [31].
However, despite these materials issues, films of (Ga,Mn)As and related com-
pounds, optimized by appropriate epitaxy and annealing procedures, show excellent
micromagnetic properties, which has opened the door to demonstrating novel func-
tionalities and to verifying the understanding of carrier-mediated ferromagnetism
and spintronic phenomena in DFS-based devices [23, 24]. In particular, it has
been found (by employing the independently determined exchange energy N0 β
as well as kp or tight-binding parameters) that the p–d Zener model (sections
“p–d Zener Model” and “p-Type Dilute Ferromagnetic Semiconductors”) even
within MFA and neglecting disorder allows one to describe semiquantitatively and
often quantitatively pertinent phenomena as a function of temperature, magnetic
field, pressure, strain, and dimensionality [23, 24]. The studied properties include
(i) thermodynamic characteristics such as Curie temperature, spin magnetization,
specific heat, carrier spin polarization, and orbital magnetization; (ii) micromagnetic
parameters including magnetic anisotropy, exchange stiffness, spin wave dispersion,
domain-wall width and intrinsic resistance; and (iii) spintronic functionalities,
e.g., electric-field control of magnetization, spin-transfer and spin-orbit torques,
intrinsic component of the anomalous Hall effect, interlayer coupling, spin injection
efficiency, current-induced domain-wall velocity and intrinsic pinning, MCD, TMR,
and TAMR. The progress in semiconductor spintronics of p-type DFSs has triggered
search for similar effects in metals that by supporting ferromagnetic ordering above
room temperature are more suitable for applications.
As an example, Fig. 9 presents the predicted and observed values of TC in p-type
(III,Mn)V compounds. As seen, there is a good agreement in TC values and chemical
trends. An analysis of theoretical results indicates that a contribution of interband
terms involving valence band subbands (the Bloembergen-Rowland mechanism) to
the magnitudes of TC is about 30%.
For verification of how TC varies with the hole concentration p, particularly
useful are studies of TC (p) in a single sample, as such a dependence is virtually
independent of poorly known values of Θp in Eq. 19 and background concentrations
of compensating donors. According to numerical results for the p–d Zener model
960 A. Bonanni et al.

Fig. 9 Magnitudes of the


Curie temperature TC
predicted within the p–d
Zener model for various
p-type III-V semiconductors
containing 5% of Mn and
3.5 × 1020 holes per cm3
(Eq. 20) compared to the
highest experimentally
determined values (upper and
lower panel, respectively).
(The data collected by
Dietl [138])

[55], γ = d ln TC /d ln p = 0.6–0.8 in the relevant region of hole densities. This


prediction was confirmed experimentally by tracing the dependence TC (p) in
(Ga,Mn)As [139] and (Ga,Mn)P [140] films irradiated by ions that produce hole-
compensating donor defects.
However, detailed studies of changes in TC induced by the gate voltage Vg in
metal-insulator-semiconductor (MIS) structures of (Ga,Mn)As [115, 141] led to an
entirely different value, γ = d ln TC /d ln(−Vg ) = 0.19 ± 0.02 [141]. As shown in
Fig. 10, this finding was elucidated by the p–d Zener model generalized to the case
of a nonuniform hole distribution obtained by solving the Poisson equation in thin
(Ga,Mn)As layers (Eq. 23), in which the Fermi level at the surface is pinned in the
gap region by surface states [115, 141].

Dilute Ferromagnetic Insulators and Topological Insulators

In the course of the years, Ga1−x Mnx N has gained the status of a model dilute
magnetic insulator, in which strong p–d coupling leads to tight binding of the hole at
a Mn acceptor (the Zhang-Rice polaron) precluding the insulator-to-metal transition,
18 Dilute Magnetic Materials 961

88 insulator (Ga,Mn)As GaAs

p(z) (10 cm )
86 1.5 -7 MV/cm

-3
84
20
1.0
82
0.5
TC (K)

3D p-d Zener
80 +7 MV/cm (uniform distribution)
0.0
78 2 nm z
p-d Zener +
76 Experiment non-uniform hole
(Device B2) distribution
74 0.23
TC v p
0.21
TC v p
72
0.5 1 1.5 2 2.5
20 -3
pS / t (10 cm )
Fig. 10 Curie temperature TC in a metal-oxide-semiconductor structure, in which the gate electric
field EG changes the hole distribution (inset) and density (the areal hole concentration ps
normalized to the thickness t of the (Ga,Mn)As channel). The solid line represents the generalized
p-d Zener model for thin layers (Eq. 23), whereas the dotted line shows the dependence predicted
by the p–d Zener model for the 3D case (Eq. 20). (From Nishitani et al. [141])

at least, up to x  0.1. Magnetic properties of Ga1−x Mnx N depend strongly on the


growth method and growth conditions [3]. In particular, spin-spin interactions are
AF in samples containing a high concentration of Mn2+ ions, presumably due to the
presence of compensating donor defects or impurities, such as Si or O. The strength
of this AF interaction is similar to that observed in Mn-based II-VI DMSs , and the
coupling is to be assigned to AF superexchange discussed in section “DMSs with
Transition Metals”.
In weakly compensated samples, where – according to intra-center optical
transitions, EPR, XANES, and XES – Mn3+ ions prevail, low-temperature ferro-
magnetism was found with the TC magnitudes reaching about 13 K at x ≈ 10%, as
shown in Fig. 11. The tight-binding model of the superexchange, which described
successfully AF interaction specific to Mn2+ ions in II-VI DMSs [73], was used
to compute exchange integrals Jij for Mn3+ pairs occupying 20 subsequent cation
neighbor positions in GaN [33]. These values of Jij incorporated to Monte Carlo
simulations provide the magnitudes of TC in agreement with experimental results
(Fig. 11).
The dependence TC (x) can also be obtained from the percolation theory,
according to which overlapping sheers start to form a percolation cluster if their
diameter rij ≥ 0.87(N0 x)−1/3 and spins are coupled if |Jij |  kB T [34].
Since the magnitudes of computed Jij decay approximately exponentially, Jij =
962 A. Bonanni et al.

1 0.125 x 0.0156 0.008


1000
experiment
theory - superexchange TBA
+ Monte Carlo
100
Curie temperature (K)

Ga1-xMnxN

10

0.1

1 2 3 4 5
x-1/3
Fig. 11 Experimental Curie temperatures TC as a function of x −1/3 in Ga1 − xMnx N (symbols)
as collected by [33] compared to theory within the tight-binding approximation and Monte Carlo
simulations, serving to determine TC (stars [33]). The slope of the straight solid line agrees
quantitatively with the percolation theory [34] for the magnitude of the exponential decay of spin-
spin coupling with the pair distance evaluated within the tight-binding approximation

J0 exp(−rij /b), one obtains [34] TC (x) = T0 exp[−0.87(N0 x)−1/3 /b], where b =
0.11 nm in Ga1−x Mnx N, the expectations corroborated by experimental and Monte
Carlo data (Fig. 11). This model implies also that all spins join the percolation
cluster only in the limit T → 0, whereas ferromagnetic and superparamagnetic
regions coexist at T > 0 in dilute ferromagnets with a finite range of spin-spin
interactions, | drJ (r)| < ∞.
An additional mechanism of FM coupling operates in topological materials. One
of the consequences of the inverted band structure specific to these systems is a
large contribution of anion p-type wave functions to Kohn-Luttinger amplitudes uik
of both the valence and conduction bands. Accordingly, appreciable magnitudes
of TF , described in Eqs. 20 and 21, can be expected from interband polarization
(the Bloembergen-Rowland mechanism) even if the Fermi energy resides in the
bandgap [51].
This FM interaction competes with AF superexchange in the case of Mn2+
ions and, as shown in Fig. 7, enhances the spin-glass freezing temperature of
Hg1−x Mnx Te, Hg1−x Mnx Se, and (Cd1−x Mnx )3 As2 in the topologically nontrivial
regime. Without strain, Hg1−x Mnx Te and Hg1−x Mnx Se are topological semimetals
for x  0.06, whereas under tensile and compressive strain, they become topolog-
ical insulators [142] and Weyl semimetals [53], respectively, in this low x limit.
18 Dilute Magnetic Materials 963

[Cr0.11(BixSb1-x)0.89]2Te3

Fig. 12 Experimental Curie temperatures TC as a function of the Bi concentration x as well as


electron and hole carrier density in a thin film of [Cr0.11 (Bix Sb1−x )0.89 ]2 Te3 . (From Chang et al.
[143])

Non-cubic (Cd1−x Mnx )3 As2 , whose band structure is similar to compressively


strained Hg1−x Mnx Te, is a Weyl semimetal over a wide Mn concentration range.
The FM Bloembergen-Rowland mechanism works together with the superex-
change in the case of early TM ions, for which the superexchange is theoretically
expected to be FM for II-VI DMSs [73] as well as for V2 -VI3 DMSs, such as
Bi2 Se3 , Bi2 Te3 , and Sb2 Se3 doped with Ti, V, or Cr [56]. Relatively high TC values
observed in (Sb,Cr)2 Te3 [144] and particularly in (Sb,V)2 Te3 [145] confirm this
view. A negligible role of bulk and boundary carriers in mediating FM coupling
is further confirmed by results for [Cr0.11 (Bix Sb1−x )0.89 ]2 Te3 presented in Fig. 12,
which demonstrate that TC is virtually independent of the Fermi-level position in
the bands and across the bandgap filled up by topological boundary states [143].
The most striking development is the demonstration of accurate quantization of
the Hall conductance |σxy | = e2 / h in thin films ferromagnetic V2 -VI3 topological
insulators in the absence of an external magnetic field under conditions that charge
transport proceeds only via the chiral edge state [50]. The precision of quantization
improves with lowering temperature, reaching the level of 10−4 in the tens of
mK range. The existence of superparamagnetic clusters in dilute ferromagnets, as
mentioned above, accounts presumably for the presence of residual backscattering
at T > 0.
964 A. Bonanni et al.

Heterogenous Magnetic Semiconductors and Oxides

The magnetic properties discussed above have concerned systems with a random
distribution of cation-substitutional magnetic impurities, except for possible Jahn-
Teller distortion for ions with a nonzero value of the orbital momentum, like the case
of Mn3+ in GaN. A challenge but also a resource of dilute magnetic materials is a
crucial dependence of their properties on the spatial distribution of magnetic ions
and their position in the crystal lattice. These nanoscale structural characteristics
depend, in turn, on the growth and processing protocols, as in the extensively studied
field of magnetic metallic alloys and nanocomposites [146] ( Chap. 15, “Metallic
Magnetic Materials”). By employing a range of photon, electron, and particle beam
methods, with structural, chemical, and spin resolution down to the nanoscale, it
has become possible to correlate surprising magnetic properties with the spatial
arrangement and the electronic configuration of the magnetic constituent [3].
A number of striking properties of nonmetallic dilute magnetic materials dis-
cussed in this section stem from a contribution of the p–d hybridization to the
binding energy, which results in attractive chemical forces between TM impurities.
Numerous ab initio studies of TM-doped semiconductors and oxides, by determin-
ing the pairing energy of TM cation dimers [147, 3], have revealed that such forces
exist in virtually all studied combinations of TM ions and nonmetallic hosts, except
for Mn in II-VI chalcogenides in which Mn-induced states lie far from the Fermi
level (see Fig. 1). These attractive forces can be overcompensated by Coulomb
repulsion if co-doping with electrically active dopants changes the charge state
of the magnetic impurities [148] though, in general, the effect of co-doping can
be more intricate leading, for instance, to the formation of impurity complexes or
affecting TM diffusion coefficients.

Phase Separation Effects in (Ga,Mn)As


It is instructive to take the canonical dilute ferromagnetic semiconductor
Ga1−x Mnx As with a nominal Mn content x ≈ 5% as an example and to consider a
palette of issues associated with the incorporation of Mn in GaAs [23, 3]:
Cation-substitutional randomly distributed Mn ions: A 3D atomic probe analysis
implies that at length scales larger than ∼1 nm, Mn ions are distributed randomly
if growth is performed below 300 ◦ C by employing low-temperature molecular
beam epitaxy (LT-MBE). These ions account for the ferromagnetic properties
of (Ga,Mn)As. However, the magnitude of the saturation magnetization points
systematically to xeff < x. An appropriate Ga/As ratio during the growth reduces
substantially the concentration of As antisite defects that act as donors compensating
the holes introduced by Mn.
Interstitial Mn ions: According to combined Rutherford backscattering (RBS)
and particle-induced x-ray emission (PIXE) measurements [149, 150], in samples
obtained by LT-MBE, a significant fraction of Mn ions occupies interstitial positions
in which they are antiferromagnetically coupled to substitutional Mn ions and form
double donors compensating the holes mediating ferromagnetic coupling. Actually,
18 Dilute Magnetic Materials 965

Fig. 13 Temperature dependence of remanent magnetization and resistivity (inset) for three
Ga0.76 Al0.24 As/Ga1−x Mnx As/Ga0.76 Al0.24 As quantum well (QW) structures. The width of the
QW is 5.6 nm, x = 0.06. Beryllium acceptors were introduced either into the first barrier (grown
before the ferromagnetic QW, I-MDH) or into the second barrier (N-MDH), or the sample was
undoped, as marked. The presence of holes originating from Be either enhances TC (N-MDH) or
reduces it by generation of Mn interstitials during the epitaxy (I-MDH). (From Wojtowicz et al.
[150])

this self-compensation effect makes the formation of the interstitials energetically


favorable, as holes in the bonding states (valence band) enlarge the system energy.
For this reason, an increase of TC by modulation doping is only effective if the
p-type barrier is deposited after the (Ga,Mn)As channel, as shown in Fig. 13.
The interstitials can be driven toward the surface and neutralized by oxidation
under low-temperature annealing Tan  200 ◦ C, which improves the ferromagnetic
characteristics.
Mn dimers: In the bulk zinc-blende structure, the crystallographic directions
[110] and [1̄10] are equivalent, whereas on the (001) surface, they are not. One
of the consequences is a different energy of NN Mn dimers oriented along these
two nonequivalent surface axes, the lower value corresponding to the [1̄10] case, as
there is an As atom below the surface connecting the two Mn ions, increasing their
binding energy due to p–d hybridization. This means that epitaxial growth results
in a surplus of [1̄10]-oriented Mn dimers. This growth-related lowering of crystal
symmetry is thought to account for the in-plane uniaxial magnetic anisotropy of
(Ga,Mn)As [151]. The existence of this anisotropy is essential for a number of
(Ga,Mn)As functionalities.
Spinodal nanodecomposition: Epitaxial growth above 300 ◦ C or annealing above
400 ◦ C of films prepared by LT MBE results, according to TEM studies, in the
966 A. Bonanni et al.

aggregation of Mn-rich (Mn,Ga)As zinc-blende nanocrystals embedded in Mn-


poor GaAs, indicating that surface and volume diffusion barriers can be overcome
at these temperatures, respectively. Though it is unknown whether (Mn,Ga)As
is on the metal or insulator side of the Mott-Hubbard transition, the magnetic
nanocrystals formed in this way are referred to as condensed magnetic semi-
conductors (CMSs). Under specific growth conditions (low Mn concentrations),
(Ga,Mn)As decomposed by the chemical phase separation shows superparamag-
netic characteristics. More often, however, superferromagnetic properties ( Chap.
20, “Magnetic Nanoparticles”) are observed. In particular, despite rather small
nanocrystal diameters, typically below 6 nm, weakly temperature-dependent open
magnetization loops persist up to 360 K [152]. This points to a rather high TC value
of the individual nanocrystals, and also to the presence of coupling between them,
perhaps mediated by a combination of strain and spin-orbit interactions [153]. In
contrast to (Ga,Mn)As grown by LT MBE, this segregated system shows many
appealing functionalities, such as large magnetooptical effects, persisting up to room
temperature [154].
MnAs precipitation: At high growth or annealing temperatures Tan  600 ◦ C,
CMSs in the form of hexagonal MnAs nanocrystals embedded coherently in zinc-
blende GaAs (crystallographic phase separation) with diameters up to 500 nm are
formed, showing magnetic properties and TC  318 K, specific to free-standing
MnAs.

Phase Separation Effects Beyond (Ga,Mn)As


There are other relevant properties associated with the low solubility of TM ions in
semiconductors and oxides.
Precipitation of TM-rich compounds or TM grains: Examples here are metallic
precipitates of Fe3 N in (Ga,Fe)N [155] and of Co in (Zn,Co)O [156], as revealed by
TEM and x-ray diffraction (XRD). They account for ferromagnetic-like signatures
in magnetic measurements persisting above room temperature and coexisting at low
temperatures with a paramagnetic component originating from diluted TM ions.
Effect of co-doping by electrically active impurities: In the above two cases, co-
doping by Si and Al donors, respectively, inhibits the precipitation, reducing the
high-temperature ferromagnetic-like response. By contrast, according to energy-
dispersive x-ray spectroscopy (EDS), co-doping with nitrogen acceptors stops the
phase separation in (Zn,Cr)Te, whereas co-doping with iodine donors, presumably
by compensating native acceptor defects, leads to an enhanced aggregation of Cr
cations, which results in room-temperature ferromagnetic-like magnetic, magne-
tooptical, and magnetotransport properties [157].
Impurity complexes: The application of extended x-ray absorption fine structure
(EXAFS) and x-ray emission spectroscopy (XES) demonstrated that co-doping of
(Ga,Mn)N with Mg acceptors results in the formation of cation complexes Mn-
Mgk , where the number of Mg ions 0 ≤ k ≤ 3 bound to Mn determines the
Mn charge and spin state, 3+ ≤ n ≤ 5+ and 2 ≥ S ≥ 1, respectively
[158]. These changes in the Mn electronic configuration alter magnetic properties
and intra-impurity optical transitions. In particular, Mn-Mg3 (Mn5+ ) complexes
show strong broadband photoluminescence in the infrared, which persists to room
18 Dilute Magnetic Materials 967

Pairing energy (eV)

wz-GaN:TM

Number of TMs in the cluster


(n)
Fig. 14 Computed pairing energies Ed as a function of the number n of TM cations in Cr, Mn,
and Fe clusters in bulk wz-GaN (circles) and on Ga(0001) surface (triangles). According to these
data, Mn ions will not, whereas Fe ions will aggregate during the epitaxy. (Adapted from Gonzalez
Szwacki et al. [159])

temperature, opening the door for laser applications in spectral windows relevant
for telecommunication [158].
Surface aggregation: Epitaxial growth proceeds under conditions allowing for
surface migration, which leads to aggregation of TM cations if the dimer pairing
energy Ed is negative at the growth surface rather than in the bulk. According to ab
initio data collected in Fig. 14, the values of Ed for Cr, Mn, and Fe cations in the bulk
GaN are negative, and |Ed | is much larger than kB T under conditions of epitaxial
growth of GaN. However, Ed becomes positive for Mn surface cation dimers. This
may explain why the solubility limit of Mn compared to that of Fe is about one order
of magnitude greater under the same growth conditions [159].
Formation of nanocolumns: There is a steady progress in controlling spinodal
nanodecomposition in DMSs, particularly the nanocrystal size distribution as well
as lateral and vertical ordering [47]. Remarkably, under specific growth conditions,
the surface aggregation and associated mismatch strain lead in some systems to the
self-organized growth of TM-rich nanocolumns embedded in the TM-poor host, an
effect observed in (Al,Cr)N [161], (Ge,Mn), and (Zn,Cr)Te [3]. The resulting TM
distribution is illustrated in Fig. 15 for the case of (Ge,Mn). These nanocolumns are
typically metallic and thus can serve as nanocontacts (in, e.g., energy harvesting
systems) or as active devices [3]. For instance, by interrupting the TM flow for a
certain time during the epitaxy, tunnel barriers can be formed resulting in a dense
array of nanoscale TMR junctions or single-electron transistors [162].
968 A. Bonanni et al.

33x33x10 nm3

28x28x50 nm3

Fig. 15 Visualization of Mn-rich nanocolumns in a Ge0.94 Mn0.06 film by 3D atomic probe. In the
nanocolumns, Ge1−x Mnx with x  0.5 is formed. (Reproduced from Mouton et al. [160], with
the permission of AIP Publishing)

Acknowledgments The work of A. B. was supported by the Austrian Science Foundation, FWF
(P31423 and P26830), and by the Austrian Exchange Service (ÖAD) project PL-01/2017. T. D.
acknowledges a support by the Foundation for Polish Science through the IRA Programme
financed by EU within SG OP Programme.

References
1. Sorokin, E., Naumov, S., Sorokina, I.T.: Ultrabroadband infrared solid-state lasers. IEEE J.
Sel. Top. Quant. Electr. 11, 690–712 (2005). https://doi.org/10.1109/JSTQE.2003.850255
2. Bonanni, A.: (Nano)characterization of semiconductor materials and structures. Semicon. Sci.
Technol. 26, 060301 (2011). https://doi.org/10.1088/0268-1242/26/6/060301
3. Dietl, T., Sato, K., Fukushima, T., Bonanni, A., Jamet, M., Barski, A., Kuroda, S., Tanaka,
M., Hai, P.N., Katayama-Yoshida, H.: Spinodal nanodecomposition in semiconductors doped
with transition metals. Rev. Mod. Phys. 87, 1311–1377 (2015). https://doi.org/10.1103/
RevModPhys.87.1311
4. Abraham, D.W., Frank, M.M., Guha, S.: Absence of magnetism in hafnium oxide films. Appl.
Phys. Lett. 87, 252502 (2005). https://doi.org/10.1063/1.2146057
5. Grace, P.J., Venkatesan, M., Alaria, J., Coey, J.M.D., Kopnov, G., Naaman, R.: The origin
of the magnetism of etched Silicon. Adv. Mater. 21, 71 (2009). https://doi.org/10.1002/adma.
200801098
6. Matsubayashi, K., Maki, M., Tsuzuki, T., Nishioka, T., Sato, N.K.: Parasitic ferromagnetism
in a hexaboride? Nature 420, 143 (2002). https://doi.org/10.1038/420143b
7. Makarova, T.L., Sundqvist, B., Höhne, R., Esquinazi, P., Kopelevich, Y., Scharff, P., Davydov,
V., Kashevarova, L.S., Rakhmanina, A.V.: Retraction: magnetic carbon. Nature 440, 707
(2006). https://doi.org/10.1038/nature04622
18 Dilute Magnetic Materials 969

8. Sawicki, M., Stefanowicz, W., Ney, A.: Sensitive SQUID magnetometry for studying
nanomagnetism. Semicon. Sci. Technol. 26, 064006 (2011). https://doi.org/10.1088/0268-
1242/26/6/064006
9. Coey, J.M.D., Venkatesan, M., Stamenov, P.: Surface magnetism of strontium titanate. J.
Phys.: Condens. Matter 28, 485001 (2016). https://doi.org/10.1088/0953-8984/28/48/485001
10. Mydosh, J.A.: Spin glasses: redux: an updated experimental/materials survey. Rep. Progr.
Phys. 78, 052501 (2015). https://doi.org/10.1088/0034-4885/78/5/052501
11. Binder, K., Young, A.P.: Spin glasses: experimental facts, theoretical concepts, and open
questions. Rev. Mod. Phys. 58, 801–976 (1986). https://doi.org/10.1103/RevModPhys.58.801
12. Kawamura, H., Taniguchi, T.: Spin glasses. In: Buschow, K.H.J. (ed.) Handbook of Magnetic
Materials, vol. 24, pp. 1–138. Elsevier, Amsterdam (2015). https://doi.org/10.0116/bs.hmm.
2015.08001
13. Stein, D.L., Newman, C.M.: Spin Glasses and Complexity. Princeton University Press,
Princeton (2013). http://press.princeton.edu/titles/9917.html
14. Rønnow, T.F., Wang, Z., Job, J., Boixo, S., Isakov, S.V., Wecker, D., Martinis, J.M., Lidar,
D.A., Troyer, M.: Defining and detecting quantum speedup. Science 345, 420–424 (2014).
https://doi.org/10.1126/science.1252319
15. Mezard, M., Parisi, G., Virasoro, M.A.: Spin glass theory and beyond. World Scientific,
Singapore (1987). https://doi.org/10.1142/0271
16. Fisher, D.S., Huse, D.A.: Equilibrium behavior of the spin-glass ordered phase. Phys. Rev. B
38, 386–411 (1988). https://doi.org/10.1103/PhysRevB.38.386
17. de Vegvar, P., Lévy, L.: Reply to comment on Conductance fluctuations of mesoscopic spin
glasses. Phys. Rev. Lett. 68, 3485–3485 (1992). https://doi.org/10.1103/PhysRevLett.68.3485
18. Bouchiat, H., Dartyge, E., Monod, P., Lambert, M.: Diffuse x-ray determination of local
atomic order in a spin-glass alloy system. Phys. Rev. B 23, 1375–1383 (1981). https://doi.
org/10.1103/PhysRevB.23.1375
19. Dietl, T., Ohno, H., Matsukura, F., Cibert, J., Ferrand, D.: Zener model description of
ferromagnetism in zinc-blende magnetic semiconductors. Science 287, 1019 (2000). https://
doi.org/10.1126/science.287.5455.1019
20. Ohno, H.: Making nonmagnetic semiconductors ferromagnetic. Science 281, 951 (1998).
https://doi.org/10.1126/science.281.5379.951
21. Story, T., Gała̧zka, R.R., Frankel, R.B., Wolff, P.A.: Carrier-concentration–induced ferromag-
netism in PbSnMnTe. Phys. Rev. Lett. 56, 777 (1986). https://doi.org/10.1103/PhysRevLett.
56.777
22. Haury, A., Wasiela, A., Arnoult, A., Cibert, J., Tatarenko, S., Dietl, T., d’Aubigne,
Y.M.: Observation of a ferromagnetic transition induced by two-dimensional hole gas in
modulation-doped CdMnTe quantum wells. Phys. Rev. Lett. 79, 511 (1997). https://doi.org/
10.1103/PhysRevLett.79.511
23. Dietl, T., Ohno, H.: Dilute ferromagnetic semiconductors: physics and spintronic structures.
Rev. Mod. Phys. 86, 187–251 (2014). https://doi.org/10.1103/RevModPhys.86.187
24. Jungwirth, T., Wunderlich, J., Novák, V., Olejník, K., Gallagher, B.L., Campion, R.P.,
Edmonds, K.W., Rushforth, A.W., Ferguson, A.J., Němec, P.: Spin-dependent phenomena
and device concepts explored in (Ga,Mn)As. Rev. Mod. Phys. 86, 855–896 (2014). https://
doi.org/10.1103/RevModPhys.86.855
25. Olejník, K., Owen, M.H.S., Novák, V., Mašek, J., Irvine, A.C., Wunderlich, J., Jungwirth, T.:
Enhanced annealing, high Curie temperature and low-voltage gating in (Ga,Mn)As: a surface
oxide control study. Phys. Rev. B 78, 054403 (2008). https://doi.org/10.1103/PhysRevB.78.
054403
26. Wang, M., Campion, R.P., Rushforth, A.W., Edmonds, K.W., Foxon, C.T., Gallagher, B.L.:
Achieving high Curie temperature in (Ga,Mn)As. Appl. Phys. Lett. 93, 132103 (2008). https://
doi.org/10.1063/1.2992200
27. Chen, L., Yang, X., Yang, F., Zhao, J., Misuraca, J., Xiong, P., von Molnar̀, S.: Enhancing the
Curie temperature of ferromagnetic semiconductor (Ga,Mn)As to 200 K via nanostructure
engineering. Nano Lett. 11, 2584 (2011). https://doi.org/10.1021/nl201187m
970 A. Bonanni et al.

28. Fukuma, Y., Asada, H., Yamamoto, J., Odawara, F., Koyanagi, T.: Large magnetic circular
dichroism of Co clusters in Co-doped ZnO. Appl. Phys. Lett. 93, 142510 (2008)
29. Hassan, M., Springholz, G., Lechner, R.T., Groiss, H., Kirchschlager, R., Bauer, G.: Molecu-
lar beam epitaxy of single phase GeMnTe with high ferromagnetic transition temperature. J.
Cryst. Growth 323, 363 (2011). https://doi.org/10.1016/j.jcrysgro.2010.10.135
30. Zhao, K., Chen, B., Zhao, G., Yuan, Z., Liu, Q., Deng, Z., Zhu, J., Jin, C.: Ferromagnetism
at 230 K in (Ba0.7 K0.3 )(Zn0.85 Mn0.15 )2 As2 diluted magnetic semiconductor. Chin. Sci. Bull.
59, 2524–2527 (2014). https://doi.org/10.1007/s11434-014-0398-z
31. Dietl, T.: Interplay between carrier localization and magnetism in diluted magnetic and
ferromagnetic semiconductors. J. Phys. Soc. Jpn. 77, 031005 (2008). https://doi.org/10.1143/
JPSJ.77.031005
32. Richardella, A., Roushan, P., Mack, S., Zhou, B., Huse, D.A., Awschalom, D.D., Yazdani, A.:
Visualizing critical correlations near the metal-insulator transition in Ga1−x Mnx As. Science
327, 665 (2010). https://doi.org/10.1126/science.1183640
33. Stefanowicz, S., Kunert, G., Simserides, C., Majewski, J.A., Stefanowicz, W., Kruse, C.,
Figge, S., Li, T., Jakieła, R., Trohidou, K.N., Bonanni, A., Hommel, D., Sawicki, M., Dietl,
T.: Phase diagram and critical behavior of a random ferromagnet Ga1−x Mnx N. Phys. Rev. B
88, 081201(R) (2013). https://doi.org/10.1103/PhysRevB.88.081201
34. Korenblit, I.Y., Shender, E.F., Shklovsky, B.I.: Percolation approach to the phase transition
in the very dilute ferromagnetic alloys. Phys. Lett. A 46, 275 (1973). https://doi.org/10.1016/
0375-9601(73)90219-3
35. Methfessel, S., Mattis, D.C.: Magnetic semiconductors. In: Flugge, S. (eds.) Handbuch der
Physik, vol. 18, p. Pt. 1. Springer, Berlin (1968)
36. Kioseoglou, G., Hanbicki, A.T., Sullivan, J.M., van ’t Erve, O.M.J., Li, C.H., Erwin, S.C.,
Mallory, R., Yasar, M., Petrou, A., Jonker, B.T.: Electrical spin injection from an n-type
ferromagnetic semiconductor into a III-V device heterostructure. Nat. Mater. 3, 799–803 (
2004). https://doi.org/10.1038/nmat1239
37. Wachter, P.: Europium chalcogenides: EuO, EuS, EuSe and EuTe. In: Gschneidner, K.A. Jr.,
Eyring, L. (eds.) Handbook on the Physics and Chemistry of Rare Earth, vol. 2, p. 507. North-
Holland, Amsterdam (1979). https://doi.org/10.1016/S0168-1273(79)02010-9
38. Nagaev, E.L.: Colossal-magnetoresistance materials: manganites and conventional fer-
romagnetic semiconductors. Phys. Rep. 346, 387 (2001). https://doi.org/10.1016/S0370-
1573(00)00111-3
39. Andrearczyk, T., Jaroszyński, J., Sawicki, M., Van Khoi, L., Dietl, T., Ferrand, D., Bour-
gognon, C., Cibert, J., Tatarenko, S., Fukumura, T., Jin, Z., Koinuma, H., Kawasaki, M.:
Ferromagnetic interactions in p- and n-type II-VI diluted magnetic semiconductors. In: Miura,
N., Ando, T. (eds.) Proceedings 25th International Conference on Physics of Semiconductors,
Osaka, 2000, p. 235. Springer, Berlin, (2001)
40. Gała̧zka, R.R.: Semimagnetic semiconductors. In: Wilson, B.L.H. (ed.) Proceedings 14th
International Conference on the Physics of Semiconductors, Edinburgh 1978, p. 133. IoP,
Bristol (1978)
41. Furdyna, J.K., Kossut, J.: Diluted Magnetic Semiconductors. Semiconductors and Semimet-
als, vol. 25. Academic Press, New York (1988)
42. Dietl, T.: (Diluted) magnetic semiconductors. In: Mahajan, S. (ed.) Handbook of Semicon-
ductors, vol. 3B, p. 1251. North Holland, Amsterdam (1994)
43. Gaj, J.A., Kossut, J. (eds.): Introduction to the Physics of Diluted Magnetic Semiconductors.
Springer, Berlin (2010). https://doi.org/10.1007/978-3-642-15856-8
44. Fiederling, R., Keim, M., Reuscher, G., Ossau, W., Schmidt, G., Waag, A., Molenkamp, L.W.:
Injection and detection of a spin-polarized current in a light-emitting diode. Nature 402, 787
(1999). https://doi.org/10.1038/45502
45. Leclercq, B., Rigaux, C., Mycielski, A., Menant, M.: Critical dynamics in Cd1−x Mnx Te spin
glasses. Phys. Rev. B 47, 6169–6172 (1993). https://doi.org/10.1103/PhysRevB.47.6169
46. Jaroszyński, J., Wróbel, J., Karczewski, G., Wojtowicz, T., Dietl, T.: Magnetoconductance
noise and irreversibilities in submicron wires of spin-glass n+ -Cd1−x Mnx Te. Phys. Rev. Lett.
80, 5635 (1998). https://doi.org/10.1103/PhysRevLett.80.5635
18 Dilute Magnetic Materials 971

47. Dietl, T.: Spin dynamics of a confined electron interacting with magnetic or nuclear spins: a
semiclassical approach. Phys. Rev. B 91, 125204 (2015). https://doi.org/10.1103/PhysRevB.
91.125204
48. Jungwirth, T., Marti, X., Wadley, P., Wunderlich, J.: Antiferromagnetic spintronics. Nat.
Nanotech. 11, 231–241 (2016). https://doi.org/10.1038/nnano.2016.18
49. Kobak, J., Smoleński, T., Goryca, M., Papaj, M., Bogucki, A., Gietka, K., Koperski, M.,
Rousset, J.-G., Suffczyński, J., Janik, E., Nawrocki, M., Golnik, A., Kossacki, P., Pacuski,
W.: Designing quantum dots for solotronics. Nat. Commun. 5, 3191 (2014). https://doi.org/
10.1038/ncomms4191
50. Chang, C.-Z., Zhang, J., Feng, X., Shen, J., Zhang, Z., Guo, M., Li, K., Ou, Y., Wei, P., Wang,
L.-L., Ji, Z.-Q., Feng, Y., Ji, S., Chen, X., Jia, J., Dai, X., Fang, Z., Zhang, S.-C., He, K., Wang,
Y., Lu, L., Ma, X.-C., Xue, Q.-K.: Experimental observation of the quantum anomalous Hall
effect in a magnetic topological insulator. Science 340, 167–170 (2013). https://doi.org/10.
1126/science.1234414
51. Yu, R., Zhang, W., Zhang, H.-J., Zhang, S.-C., Dai, X., Fang, Z.: Quantized anomalous Hall
effect in magnetic topological insulators. Science 329, 61–64 (2010). https://doi.org/10.1126/
science.1187485
52. Fan, Y., Kou, X., Upadhyaya, P., Shao, Q., Pan, L., Lang, M., Che, X., Tang, J., Montazeri,
M., Murata, K., et al.: Electric-field control of spin–orbit torque in a magnetically doped
topological insulator. Nat. Nanotechnol. 352 (2016). https://doi.org/10.1038/nnano.2015.294
53. Bulmash, D., Liu, C.-X., Qi, X.-L.: Prediction of a Weyl semimetal in Hg1−x−y Cdx Mny Te.
Phys. Rev. B 89, 081106 (2014). https://doi.org/10.1103/PhysRevB.89.081106
54. Gała̧zka, R.R.: II–VI compounds – Polish perspective. Phys. Stat. Sol. (B) 243, 759 (2006).
https://doi.org/10.1002/pssb.200564601
55. Dietl, T., Ohno, H., Matsukura, F.: Hole-mediated ferromagnetism in tetrahedrally coordi-
nated semiconductors. Phys. Rev. B 63, 195205 (2001). https://doi.org/10.1103/PhysRevB.
63.195205
56. Vergniory, M.G., Otrokov, M.M., Thonig, D., Hoffmann, M., Maznichenko, I.V., Geilhufe,
M., Zubizarreta, X., Ostanin, S., Marmodoro, A., Henk, J., Hergert, W., Mertig, I., Chulkov,
E.V., Ernst, A.: Exchange interaction and its tuning in magnetic binary chalcogenides. Phys.
Rev. B 89, 165202 (2014). https://doi.org/10.1103/PhysRevB.89.165202
57. Bonanni, A., Dietl, T.: A story of high-temperature ferromagnetism in semiconductors. Chem.
Soc. Rev. 39, 528 (2010). https://doi.org/10.1039/b905352m
58. Fukushima, T., Sato, K., Katayama-Yoshida, H., Dederichs, P.H.: Spinodal decomposition
under layer by layer growth condition and high Curie temperature quasi-one-dimensional
nano-structure in dilute magnetic semiconductors. Jpn. J. Appl. Phys. 45, L416 (2006). https://
doi.org/10.1143/JJAP.45.L416
59. Ludwig, G.W., Woodbury, H.H.: Electron spin resonance in semiconductors, pp. 223–304.
Academic Press, New York (1962)
60. Zunger, A.: Electronic structure of 3d transition-atom impurities in semiconductors. In: Seitz,
F., Turnbull, D. (eds.) Solid State Physics, vol. 39, p. 275. Academic Press, New York
(1986)
61. Langer, J.M., Delerue, C., Lannoo, M., Heinrich, H.: Transition-metal impurities in semicon-
ductors and heterojunction band lineups. Phys. Rev. B 38, 7723 (1988). https://doi.org/10.
1103/PhysRevB.38.7723
62. Nolte, D.D.: Semi-insulating semiconductor heterostructures: optoelectronic properties and
applications. J. Appl. Phys. 85, 6259 (1999). https://doi.org/10.1063/1.370284
63. Khomskii, D.I.: Transition Metal Compounds. Cambridge University Press, Cambridge
(2014). https://doi.org/10.1017/CBO9781139096782
64. Pacuski, W., Suffczyński, J., Osewski, P., Kossacki, P., Golnik, A., Gaj, J.A., Deparis, C.,
Morhain, C., Chikoidze, E., Dumont, Y., Ferrand, D., Cibert, J., Dietl, T.: Influence of s,p-d
and s-p exchange couplings on exciton splitting in Zn1−x Mnx O. Phys. Rev. B 84, 035214
(2011). https://doi.org/10.1103/PhysRevB.84.035214
65. Belitz, D., Kirkpatrick, T.R.: The Anderson-Mott transition. Rev. Mod. Phys. 66, 261 (1994).
https://doi.org/10.1103/RevModPhys.66.261
972 A. Bonanni et al.

66. Byczuk, K., Hofstetter, W., Vollhardt, D.: Anderson localization vs. Mott-Hubbard metal-
insulator transition in disordered, interacting lattice fermion systems. J. Mod. Phys. B 24,
1727 (2010). https://doi.org/10.1142/S0217979210064575
67. Shklovskii, B.I., Efros, A.L.: Electronic Properties of Doped Semiconductors, Springer,
Berlin (1984). https://doi.org/10.1007/978-3-662-02403-4
68. Wilamowski, Z., Świa̧tek, K., Dietl, T., Kossut, J.: Resonant states in semiconductors: a
quantitative study of HgSe:Fe. Solid State Commun. 74, 833 (1990). https://doi.org/10.1016/
0038-1098(90)90945-8
69. Sawicki, M., Dietl, T., Plesiewicz, W., Sekowski, P., Sniadower, L., Baj, M., Dmowski, L.:
Influence of an acceptor state on transport in zero-gap Hg1−x Mnx Te. In: Landwehr, G. (ed.)
Application of High Magnetic Fields in Physics of Semiconductors, p. 382. Springer, Berlin
(1983). https://doi.org/10.1007/3-540-11996-5_55
70. Story, T., Górska, M., Łusakowski, A., Arciszewska, M., Dobrowolski, W., Grodzicka,
E., Gołacki, Z., Gała̧zka, R.R.: New mechanism of f –f exchange interactions controlled
by Fermi level position. Phys. Rev. Lett. 77, 3447–3450 (1996). https://doi.org/10.1103/
PhysRevLett.77.3447
71. Dzero, M., Xia, J., Galitski, V., Coleman, P.: Topological Kondo insulators. Ann. Rev. Con-
den. Matter Phys. 7, 249–280 (2016). https://doi.org/10.1146/annurev-conmatphys-031214-
014749
72. Dietl, T., Śliwa, C., Bauer, G., Pascher, H.: Mechanisms of exchange interactions between
carriers and Mn or Eu spins in lead chalcogenides. Phys. Rev. B 49, 2230 (1994). https://doi.
org/10.1103/PhysRevB.49.2230
73. Kacman, P.: Spin interactions in diluted magnetic semiconductors and magnetic semiconduc-
tor structures. Semicond. Sci. Technol. 16, R25 (2001). https://doi.org/10.1088/0268-1242/
16/4/201
74. Chanier, T., Virot, F., Hayn, R.: Chemical trend of exchange coupling in diluted magnetic
II-VI semiconductors: ab initio calculations. Phys. Rev. B 79, 205204 (2009). https://doi.org/
10.1103/PhysRevB.79.205204
75. Adhikari, R., Stefanowicz, W., Faina, B., Capuzzo, G., Sawicki, M., Dietl, T., Bonanni, A.:
Upper bound for the s-d exchange integral in n-(Ga,Mn)N:Si from magnetotransport studies.
Phys. Rev. B 91, 205204 (2015). https://doi.org/10.1103/PhysRevB.91.205204
76. Steeneken, P.G., Tjeng, L.H., Elfimov, I., Sawatzky, G.A., Ghiringhelli, G., Brookes, N.B.,
Huang, D.-J.: Exchange splitting and charge carrier spin polarization in EuO. Phys. Rev. Lett.
88, 047201 (2002). https://doi.org/10.1103/PhysRevLett.88.047201
77. Irkhin, V.Y., Irkhin, Y.P.: Hybridization and Kondo effect in systems with degenerate d and f
shells. Zh. Eksp. Teor. Fiz. 107, 616–627 (1995). www.jetp.ac.ru/cgi-bin/dn/e_080_02_0334.
pdf
78. Mac, W., Khoi, N.T., Twardowski, A., Gaj, J.A., Demianiuk, M.: Ferromagnetic p–d
exchange in Zn1-x Crx Se diluted magnetic semiconductor. Phys. Rev. Lett. 71, 2327–2330
(1993). https://doi.org/10.1103/PhysRevLett.71.2327
79. Winkler, R.: Spin-Orbit Coupling Effects in Two-Dimensional Electron and Hole Systems.
Springer Tracts in Modern Physics, vol. 191 Springer, Berlin (2003). Note in this book a
convention is adopted according to which a sign of the sp-d Hamiltonian is reversed. http://
www.springer.com/gp/book/9783540011873
80. Jensen, J., Mackintosh, A.R.: Rare Earth Magnetism. Structures and Excitations. The
International Series of Monographs on Physics Clarendon Press, Oxford (1991). www.fys.
ku.dk/~jjensen/Book/Ebook.pdf
81. Sankowski, P., Kacman, P., Majewski, J.A., Dietl, T.: Spin-dependent tunneling in modulated
structures of (Ga,Mn)As. Phys. Rev. B 75, 045306 (2007). https://doi.org/10.1103/PhysRevB.
75.045306
82. Qi, X.-L., Zhang, S.-C.: Topological insulators and superconductors. Rev. Mod. Phys. 83,
1057–1110 (2011). https://doi.org/10.1103/RevModPhys.83.1057
83. Bauer, G., Pascher, H., Zawadzki, W.: Magneto-optical properties of semimagnetic lead
chalcogenides. Semicond. Sci. Technol. 7, 703 (1992). https://doi.org/10.1088/0268-1242/
7/6/001
18 Dilute Magnetic Materials 973

84. Liu, C.-X., Qi, X.-L., Dai, X., Fang, Z., Zhang, S.-C.: Quantum anomalous Hall effect in
Hg1−y Mny Te quantum wells. Phys. Rev. Lett. 101, 146802 (2008). https://doi.org/10.1103/
PhysRevLett.101.146802
85. Barnes, S.E.: Theory of electron spin resonance of magnetic ions in metals. Adv. Phys. 30,
801–938 (1981). https://doi.org/10.1080/00018738100101447
86. Dietl, T.: Hole states in wide band-gap diluted magnetic semiconductors and oxides. Phys.
Rev. B 77, 085208 (2008). https://doi.org/10.1103/PhysRevB.77.085208
87. Śliwa, C., Dietl, T.: Electron-hole contribution to the apparent s–d exchange interaction in III-
V dilute magnetic semiconductors. Phys. Rev. B 78, 165205 (2008). https://doi.org/10.1103/
PhysRevB.78.165205
88. Toyoda, T., Kume, K.: Residual resistivities of Au- or Cu-based 4d transition metal
dilute alloys. Solid State Commun. 15, 1889–1890 (1974). https://doi.org/10.1016/0038-
1098(74)90110-0
89. Monod, P., Schultz, S.: Conduction electron spin-flip scattering by impurities in copper.
J. Phys. (Paris) 43, 393–401 (1982). https://doi.org/10.1051/jphys:01982004302039300
90. Braspenning, P.J., Zeller, R., Dederichs, P.H., Lodder, A.: Electronic structure of non-
magnetic impurities in Cu. J. Phys. F: Metal Phys. 12, 105 (1982). https://doi.org/10.1088/
0305-4608/12/1/011
91. Dietl, T., Sawicki, M., Dahl, M., Heiman, D., Isaacs, E.D., Graf, M.J., Gubarev, S.I., Alov,
D.L.: Spin-flip scattering near the metal-to-insulator transition in Cd0.95 Mn0.05 Se:In. Phys.
Rev. B 43, 3154–3163 (1991). https://doi.org/10.1103/PhysRevB.43.3154
92. Fisher, M.E., Langer, J.S.: Resistive anomalies at magnetic critical points. Phys. Rev. Lett.
20, 665 (1968). https://doi.org/10.1103/PhysRevLett.20.665
93. Novák, V., Olejník, K., Wunderlich, J., Cukr, M., Výborný, K., Rushforth, A.W., Edmonds,
K.W., Campion, R.P., Gallagher, B.L., Sinova, J., Jungwirth, T.: Curie point singularity in
the temperature derivative of resistivity in (Ga,Mn)As. Phys. Rev. Lett. 101, 077201 (2008).
https://doi.org/10.1103/PhysRevLett.101.077201
94. Yakovlev, D.R., Ossau, W.: Magnetic polarons. In: Gaj, J.A., Kossut, J. (eds.) Introduction
to the Physics of Diluted Magnetic Semiconductors. Springer Series in Materials Science,
vol. 144. Spinger, Berlin (2010). https://doi.org/10.1007/978-3-642-15856-8_7
95. Benoit á La Guillaume, C.: Free magnetic polarons in three, quasi-two, and quasi-one dimen-
sions. phys. stat. sol. (B) 175, 369–380 (1993). https://doi.org/10.1002/pssb.2221750208
96. Peterson, D.L., Bartholomew, D.U., Ramdas, A.K., Rodriguez, S.: Magnetic field and
temperature tuning of resonant Raman scattering in diluted magnetic semiconductors. Phys.
Rev. B 31, 7932 (1985). https://doi.org/10.1103/PhysRevB.31.7932
97. Dietl, T., Peyla, P., Grieshaber, W., Merle d’Aubigné, Y.: Dynamics of spin organization in
diluted magnetic semiconductors. Phys. Rev. Lett. 74, 474 (1995). https://doi.org/10.1103/
PhysRevLett.74.474
98. Pierre, F., Birge, N.O.: Dephasing by extremely dilute magnetic impurities revealed by
Aharonov-Bohm oscillations. Phys. Rev. Lett. 89, 206804 (2002). https://doi.org/10.1103/
PhysRevLett.89.206804
99. Jaroszyński, J., Wróbel, J., Sawicki, M., Kamińska, E., Skośkiewicz, T., Karczewski, G.,
Wojtowicz, T., Piotrowska, A., Kossut, J., Dietl, T.: Influence of s-d exchange interaction
on universal conductance fluctuations in Cd1−x Mnx Te:In. Phys. Rev. Lett. 75, 3170 (1995).
https://doi.org/10.1103/PhysRevLett.75.3170
100. Wojtowicz, T., Dietl, T., Sawicki, M., Plesiewicz, W., Jaroszyński, J.: Metal-insulator
transition in semimagnetic semiconductors. Phys. Rev. Lett. 56, 2419–2422 (1986). https://
doi.org/10.1103/PhysRevLett.56.2419
101. Nagaosa, N., Sinova, J., Onoda, S., MacDonald, A.H., Ong, N.P.: Anomalous Hall effect. Rev.
Mod. Phys. 82, 1539 (2010). https://doi.org/10.1103/RevModPhys.82.1539
102. Ralph, D.C., Stiles, M.D.: Spin transfer torques. J. Magn. Magn. Mater. 320, 1190–1216
(2008). https://doi.org/10.1016/j.jmmm.2007.12.019
103. Manchon, A., Koo, H.C., Nitta, J., Frolov, S.M., Duine, R.A.: New perspectives for Rashba
spin-orbit coupling. Nat. Mater. 14, 871–882 (2015). https://doi.org/10.1038/nmat4360
974 A. Bonanni et al.

104. Hasan, M.Z., Kane, C.L.: Colloquium: topological insulators. Rev. Mod. Phys. 82, 3045–3067
(2010). https://doi.org/10.1103/RevModPhys.82.3045
105. Barborini, M.: Neutral, anionic, and cationic manganese dimers through density functional
theory. J. Phys. Chem. A 120, 1716–1726 (2016). https://doi.org/10.1021/acs.jpca.5b12169
106. Shapira, Y., Bindilatti, V.: Magnetization-step studies of antiferromagnetic clusters and single
ions: exchange, anisotropy, and statistics. J. Appl. Phys. 92, 4155 (2002). https://doi.org/10.
1063/1.1507808
107. Savoyant, A., D’Ambrosio, S., Kuzian, R.O., Daré, A.M., Stepanov, A.: Exchange integrals
in Mn- and Co-doped II-VI semiconductors. Phys. Rev. B 90, 075205 (2014). https://doi.org/
10.1103/PhysRevB.90.075205
108. Story, T.: Semimagnetic semiconductors based on lead chalcogenides,” In: Khokhlov, D. (ed.)
Lead Chalcogenides, Physics and Applications. Optoelectronic Properties of Semiconductors
and Supelattices, vol. 18. Taylor and Francis, New York, (2002)
109. Samarth, N., Furdyna, J.K.: A proposed interpretation of EPR linewidth in diluted magnetic
semiconductors. Solid State Commun. 65, 801–804 (1988). https://doi.org/10.1016/0038-
1098(88)90508-X
110. Larson, B.E., Ehrenreich, H.: Anisotropic superexchange and spin-resonance linewidth in
diluted magnetic semiconductors. Phys. Rev. B 39, 1747–1759 (1989). https://doi.org/10.
1103/PhysRevB.39.1747
111. Blinowski, J., Kacman, P., Majewski, J.A.: Ferromagnetic superexchange in Cr-based diluted
magnetic semiconductors. Phys. Rev. B 53, 9524 (1996). https://doi.org/10.1103/PhysRevB.
53.9524
112. Söllinger, W., Heiss, W., Lechner, R.T., Rumpf, K., Granitzer, P., Krenn, H., Springholz, G.:
Exchange interactions in europium monochalcogenide magnetic semiconductors and their
dependence on hydrostatic strain. Phys. Rev. B 81, 155213 (2010). https://doi.org/10.1103/
PhysRevB.81.155213
113. Kuneš, J., Ku, W., Pickett, W.E.: Exchange coupling in Eu monochalcogenides from first
principles. J. Phys. Soc. Japan 74, 1408–1411 (2005). https://doi.org/10.1143/JPSJ.74.1408
114. Lewiner, C., Bastard, G.: Indirect exchange interaction in extremely non-parabolic zero-gap
semiconductors. J. Phys. C: Solid State Phys. 13, 2347 (1980). https://doi.org/10.1088/0022-
3719/13/12/016
115. Sawicki, M., Chiba, D., Korbecka, A., Nishitani, Y., Majewski, J.A., Matsukura, F., Dietl, T.,
Ohno, H. Experimental probing of the interplay between ferromagnetism and localization in
(Ga, Mn)As. Nat. Phys. 6, 22 (2010). https://doi.org/10.1038/nphys1455
116. Dietl, T., Haury, A., d’Aubigne, Y.M.: Free carrier-induced ferromagnetism in structures of
diluted magnetic semiconductors. Phys. Rev. B 55, R3347 (1997). https://doi.org/10.1103/
PhysRevB.55.R3347
117. Dietl, T., Cibert, J., Ferrand, D., Merle d’Aubigné, Y.: Carrier-mediated ferromagnetic
interactions in structures of magnetic semiconductors. Mater. Sci. Eng. B 63, 103–110 (1999).
https://doi.org/10.1016/S0921-5107(99)00059-8
118. Boukari, H., Kossacki, P., Bertolini, M., Ferrand, D., Cibert, J., Tatarenko, S., Wasiela,
A., Gaj, J.A., Dietl, T.: Light and electric field control of ferromagnetism in magnetic
quantum structures. Phys. Rev. Lett. 88, 207204 (2002). https://doi.org/10.1103/PhysRevLett.
88.207204
119. Liu, Q., Liu, C.-X., Xu, C., Qi, X.-L., Zhang, S.-C.: Magnetic impurities on the surface
of a topological insulator. Phys. Rev. Lett. 102, 156603 (2009). https://doi.org/10.1103/
PhysRevLett.102.156603
120. Abanin, D.A., Pesin, D.A.: Ordering of magnetic impurities and tunable electronic properties
of topological insulators. Phys. Rev. Lett. 106, 136802 (2011). https://doi.org/10.1103/
PhysRevLett.106.136802
121. Maletta, H., Felsch, W.: Insulating spin-glass system Eux Sr1−x S. Phys. Rev. B 20, 1245–1260
(1979). https://doi.org/10.1103/PhysRevB.20.1245
122. Vincent, E.: Ageing, rejuvenation and memory: the example of spin-glasses, pp. 7–60.
Springer, Berlin (2007)
18 Dilute Magnetic Materials 975

123. Andresen, J.C., Katzgraber, H.G., Oganesyan, V., Schechter, M.: Existence of a thermody-
namic spin-glass phase in the zero-concentration limit of anisotropic dipolar systems. Phys.
Rev. X 4, 041016 (2014). https://doi.org/10.1103/PhysRevX.4.041016
124. Bouchiat, H.: Determination of the critical exponents in the AgMn spin glass. J. de Phys.
(Paris) 47, 71–88 (1986). https://doi.org/10.1051/jphys:0198600470107100
125. Gunnarsson, K., Svedlindh, P., Nordblad, P., Lundgren, L., Aruga, H., Ito, A.: Dynamics of
an Ising spin-glass in the vicinity of the spin-glass temperature. Phys. Rev. Lett. 61, 754–757
(1988). https://doi.org/10.1103/PhysRevLett.61.754
126. Fert, A., De Courtenay, N., Bouchiat, H.: Experimental arguments for a non-zero transition
temperature in RKKY Heisenberg spin glasses without anisotropy. J. de Physique 49, 1173–
1178 (1988). https://doi.org/10.1051/jphys:019880049070117300
127. Larson, B.E., Ehrenreich, H.: Exchange in II-VI-based magnetic semiconductors (invited).
J. Appl. Phys. 67, 5084–5089 (1990). https://doi.org/10.1063/1.344681
128. Campos, I., Cotallo-Aban, M., Martin-Mayor, V., Perez-Gaviro, S., Taranco, A.: Spin-glass
transition of the three-dimensional Heisenberg spin glass. Phys. Rev. Lett. 97, 217204 (2006).
https://doi.org/10.1103/PhysRevLett.97.217204
129. Fabris, F.W., Pureur, P., Schaf, J., Vieira, V.N., Campbell, I.A.: Chiral anomalous Hall effect
in reentrant AuFe alloys. Phys. Rev. B 74, 214201 (2006). https://doi.org/10.1103/PhysRevB.
74.214201
130. Bouchiat, H., de Courtenay, N., Monod, P., Ocio, M., Refregier, P.: Critical behavior and
magnetic fluctuations in spin glasses. Japan. J. Appl. Phys. 26, S3-3, 1951 (1987). https://doi.
org/10.7567/JJAPS.26S3.1951
131. Weissman, M.B.: What is a spin glass? A glimpse via mesoscopic noise. Rev. Mod. Phys. 65,
829–839 (1993). https://doi.org/10.1103/RevModPhys.65.829
132. Weissman, M.B., Israeloff, N.E., Alers, G.B.: Spin-glass fluctuation statistics: mesoscopic
experiments in CuMn. J. Magn. Magn. Mater. 114, 87–130 (1992). https://doi.org/10.1016/
0304-8853(92)90336-M
133. Hor, Y.S., Roushan, P., Beidenkopf, H., Seo, J., Qu, D., Checkelsky, J.G., Wray, L.A., Hsieh,
D., Xia, Y., Xu, S.-Y., Qian, D., Hasan, M.Z., Ong, N.P., Yazdani, A., Cava, R.J.: Development
of ferromagnetism in the doped topological insulator Bi2−x Mnx Te3 . Phys. Rev. B 81, 195203
(2010). https://doi.org/10.1103/PhysRevB.81.195203
134. Park, Y.D., Hanbicki, A.T., Erwin, S.C., Hellberg, C.S., Sullivan, J.M., Mattson, J.E.,
Ambrose, T.F., Wilson, A., Spanos, G., Jonker, B.T.: A group-IV ferromagnetic semicon-
ductor: Mnx Ge1−x . Science 295, 651 (2002). https://doi.org/10.1126/science.1066348
135. Deng, Z., Jin, C.Q., Liu, Q.Q., Wang, X.C., Zhu, J.L., Feng, S.M., Chen, L.C., Yu, R.C.,
Arguello, C., Goko, T., Ning, F., Zhang, J., Wang, Y., Aczel, A.A., Munsie, T., Williams,
T.J., Luke, G.M., Kakeshita, T., Uchida, S., Higemoto, W., Ito, T.U., Gu, B., Maekawa, S.,
Morris, G.D., Uemura, Y.J.: Li(Zn,Mn)As as a new generation ferromagnet based on a I-II-V
semiconductor. Nat. Commun. 2, 422 (2011). https://doi.org/10.1038/ncomms1425
136. Przybylińska, H., Springholz, G., Lechner, R.T., Hassan, M., Wegscheider, M., Jantsch,
W., Bauer, G.: Magnetic-field-induced ferroelectric polarization reversal in the multiferroic
Ge1−x Mnx Te semiconductor. Phys. Rev. Lett. 112, 047202 (2014). https://doi.org/10.1103/
PhysRevLett.112.047202
137. Wang, K.Y., Sawicki, M., Edmonds, K.W., Campion, R.P., Rushforth, A.W., Freeman, A.A.,
Foxon, C.T., Gallagher, B.L., Dietl, T.: Control of coercivities in (Ga,Mn)As thin films by
small concentrations of mnas nanoclusters. Appl. Phys. Lett. 88, 022510 (2006). https://doi.
org/10.1063/1.2162856
138. Dietl, T.: A ten-year perspective on dilute magnetic semiconductors and oxides. Nat. Mater.
9, 965 (2010). https://doi.org/10.1038/nmat2898
139. Mayer, M.A., Stone, P.R., Miller, N., Smith, H.M., Dubon, O.D., Haller, E.E., Yu, K.M.,
Walukiewicz, W., Liu, X., Furdyna, J.K.: Electronic structure of Ga1−x Mnx As analyzed
according to hole-concentration-dependent measurements. Phys. Rev. B 81, 045205 (2010).
https://doi.org/10.1103/PhysRevB.81.045205
976 A. Bonanni et al.

140. Winkler, T.E., Stone, P.R., Li, T., Yu, K.M., Bonanni, A., Dubon, O.D.: Compensation-
dependence of magnetic and electrical properties in Ga1−x Mnx P. Appl. Phys. Lett. 98,
012103 (2011). https://doi.org/10.1063/1.3535957
141. Nishitani, Y., Chiba, D., Endo, M., Sawicki, M., Matsukura, F., Dietl, T., Ohno, H.: Curie
temperature versus hole concentration in field-effect structures of Ga1−x Mnx As. Phys. Rev.
B 81, 045208 (2010). https://doi.org/10.1103/PhysRevB.81.045208
142. Brüne, C., Liu, C.X., Novik, E.G., Hankiewicz, E.M., Buhmann, H., Chen, Y.L., Qi, X.L.,
Shen, Z.X., Zhang, S.C., Molenkamp, L.W.: Quantum Hall effect from the topological surface
states of strained bulk HgTe. Phys. Rev. Lett. 106, 126803 (2011). https://doi.org/10.1103/
PhysRevLett.106.126803
143. Chang, C.-Z., Zhang, J., Liu, M., Zhang, Z., Feng, X., Li, K., Wang, L.-L., Chen, X., Dai,
X., Fang, Z., Qi, X.-L., Zhang, S.-C., Wang, Y., He, K., Ma, X.-C., Xue, Q.-K.: Thin films of
magnetically doped topological insulator with carrier-independent long-range ferromagnetic
order. Adv. Mater. 25, 1065–1070 (2013). https://doi.org/10.1002/adma.201203493
144. Zhou, Z., Chien, Y.-J., Uher, C.: Thin film dilute ferromagnetic semiconductors Sb2−x Crx Te3
with a Curie temperature up to 190 K. Phys. Rev. B 74, 224418 (2006). https://doi.org/10.
1103/PhysRevB.74.224418
145. Zhou, Z., Chien, Y.-J., Ctirad, U.: Thin-film ferromagnetic semiconductors based on
Sb2−x Vx Te3 with TC of 177 K. Appl. Rev. Lett. 87, 112503 (2005). https://doi.org/10.1063/
1.2045561
146. Enders, A., Skomski, R., Sellmyer, D.J.: Designed magnetic nanostructures. In: Liu, J.P.,
Fullerton, E., Gutfleisch, O., Sellmyer, D.J. (eds.) Nanoscale Magnetic Materials and
Applications. Springer (2009). https://doi.org/10.1007/978-0-387-85600-1_3
147. Sato, K., Bergqvist, L., Kudrnovský, J., Dederichs, P.H., Eriksson, O., Turek, I., Sanyal,
B., Bouzerar, G., Katayama-Yoshida, H., Dinh, V.A., Fukushima, T., Kizaki, H., Zeller, R.:
First-principles theory of dilute magnetic semiconductors. Rev. Mod. Phys. 82, 1633 (2010).
https://doi.org/10.1103/RevModPhys.82.1633
148. Dietl, T.: Self-organised growth controlled by charge states of magnetic impurities. Nat.
Mater. 5, 673 (2006). https://doi.org/10.1038/nmat1721
149. Yu, K.M., Walukiewicz, W., Wojtowicz, T., Kuryliszyn, I., Liu, X., Sasaki, Y., Furdyna, J.K.:
Effect of the location of Mn sites in ferromagnetic Ga1−x Mnx As on its Curie temperature.
Phys. Rev. B 65, 201303 (2002). https://doi.org/10.1103/PhysRevB.65.201303
150. Wojtowicz, T., Lim, W.L., Liu, X., Dobrowolska, M., Furdyna, J.K., Yu, K.M., Walukiewicz,
W., Vurgaftman, I., Meyer, J.R.: Enhancement of Curie temperature in Ga1−x Mnx As/
Ga1−y Aly As ferromagnetic heterostructures by Be modulation doping. Appl. Phys. Lett. 83,
4220 (2003). https://doi.org/10.1063/1.1628815
151. Birowska, M., Śliwa, C., Majewski, J.A., Dietl, T.: Origin of bulk uniaxial anisotropy in zinc-
blende dilute magnetic semiconductors. Phys. Rev. Lett. 108, 237203 (2012). https://doi.org/
10.1103/PhysRevLett.108.237203
152. Moreno, M., Trampert, A., Jenichen, B., Däweritz, L., Ploog, K.H.: Correlation of structure
and magnetism in GaAs with embedded Mn(Ga)As magnetic nanoclusters. J. Appl. Phys. 92,
4672 (2002). https://doi.org/10.1063/1.1506402
153. Korenev, V.L., Salewski, M., Akimov, I.A., Sapega, V.F., Langer, L., Kalitukha, I.V., Debus, J.,
Dzhioev, R.I., Yakovlev, D.R., Müller, D., Schröder, C., Hövel, H., Karczewski, G., Wiater,
M., Wojtowicz, T., Kusrayev, Y.G., Bayer, M.: Long-range p–d exchange interaction in a
ferromagnet-semiconductor hybrid structure. Nat. Phys. 12, 85–91 (2016). https://doi.org/10.
1038/nphys3497
18 Dilute Magnetic Materials 977

154. Tanaka, M.: Ferromagnet (MnAs)/III-V semiconductor hybrid structures. Semicond. Sci.
Technol. 17, 327 (2002). https://doi.org/10.1088/0268-1242/17/4/306
155. Bonanni, A., Navarro-Quezada, A., Li, T., Wegscheider, M., Matĕj, Z., Holý, V., Lechner,
R.T., Bauer, G., Rovezzi, M., D’Acapito, F., Kiecana, M., Sawicki, M., Dietl, T.: Controlled
aggregation of magnetic ions in a semiconductor: an experimental demonstration. Phys. Rev.
Lett. 101, 135502 (2008). https://doi.org/10.1103/PhysRevLett.101.135502
156. Venkatesan, M., Stamenov, P., Dorneles, L.S., Gunning, R.D., Bernoux, B., Coey, J.M.D.:
Magnetic, magnetotransport, and optical properties of Al-doped Zn0.95 Co0.05 O thin films.
Appl. Phys. Lett. 90, 242508 (2007). https://doi.org/10.1063/1.2748343
157. Kuroda, S., Nishizawa, N., Takita, K., Mitome, M., Bando, Y., Osuch, K., Dietl, T.: Origin
and control of high temperature ferromagnetism in semiconductors. Nat. Mater. 6, 440 (2007).
https://doi.org/10.1038/nmat1910
158. Devillers, T., Rovezzi, M., Gonzalez Szwacki, N., Dobkowska, S., Stefanowicz, W., Sztenkiel,
D., Grois, A., Suffczyński, J., Navarro-Quezada, A., Faina, B., Li, T., Glatzel, P., d’Acapito,
F., Jakieła, R., Sawicki, M., Majewski, J.A., Dietl, T., Bonanni, A.: Manipulating Mn–Mgk
cation complexes to control the charge- and spin-state of Mn in GaN. Sci. Rep. 2, 722 (2012).
https://doi.org/10.1038/srep00722
159. Gonzalez Szwacki, N., Majewski, J.A., Dietl, T.: Aggregation and magnetism of Cr, Mn, and
Fe cations in GaN. Phys. Rev. B 83, 184417 (2011). https://doi.org/10.1103/PhysRevB.83.
184417
160. Mouton, I., Lardé, R., Talbot, E., Cadel, E., Genevois, C., Blavette, D., Baltz, V., Prestat, E.,
Guillemaud, P.B., Barski, A., Jamet, M.: Composition and morphology of self-organized Mn-
rich nanocolumns embedded in Ge: correlation with the magnetic properties. J. Appl. Phys.
112, 113918 (2012). https://doi.org/10.1063/1.4768723
161. Gu, L., Wu, S., Liu, H., Singh, R., Newman, N., Smith, D.: Characterization of Al(Cr)N and
Ga(Cr)N dilute magnetic semiconductors. J. Magn. Magn. Mater. 290, 1395 (2005). https://
doi.org/10.1016/j.jmmm.2004.11.446
162. Fukushima, T., Sato, K., Katayama-Yoshida, H.: Ab initio design of fabrication process and
shape control of self-organized tera-bit-density nano-magnets in dilute magnetic semiconduc-
tors by two-dimensional spinodal decomposition. Phys. Stat. Sol. (C) 3, 4139 (2007). https://
doi.org/10.1002/pssc.200672830

Alberta Bonanni graduated in physics at the University of


Trieste – Italy. She worked at the University of Minnesota and at
the Synchrotron Center of the University of Madison-Wisconsin
in the USA. She is now at the Johannes Kepler University in Linz,
Austria, and her scientific interests are focused on the physics of
nitride compounds and quantum materials.
978 A. Bonanni et al.

Tomasz Dietl obtained his PhD from the Institute of Physics,


Polish Academy of Sciences in Warsaw, where he is presently
a head of the International Centre for Interfacing Magnetism and
Superconductivity with Topological Matter “MagTop.” He is also
a PI at the Advanced Institute for Materials Research at Tohoku
University in Sendai, Japan.

Hideo Ohno received his PhD from the University of Tokyo in


1982. He joined Hokkaido University from 1982 and he was a
visiting scientist at the IBM T. J. Watson Research Center for
1.5 years. He was appointed professor at Tohoku University in
1994 and is president since 2018. His research interests include
spintronics and semiconductor science and technology.
Single-Molecule Magnets and Molecular
Quantum Spintronics 19
Gheorghe Taran, Edgar Bonet, and Wolfgang Wernsdorfer

Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 980
Spin Hamiltonian . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 982
Single Ion Spin Hamiltonian . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 982
Multi-spin Hamiltonian . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 987
Giant Spin Hamiltonian . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 988
Quantum Tunneling of Magnetization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 988
Landau–Zener–Stückelberg (LZS) Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 989
Spin Parity and Quantum Phase Interference . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 991
Quantum Coherence in Molecular Magnets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 994
Resonant Photon Absorption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 994
Rabi Oscillations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 994
Molecular Quantum Spintronics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 995
Direct Coupling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 996
Indirect Coupling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 999
Quantum Algorithms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1001
Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1002
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1002

Abstract

This chapter gives an overview of the main phenomenologies related to the


magnetism of single-molecule magnets (SMMs) and covers some important
achievements in the field of molecular spintronics. We start by discussing the
dominant interactions at sub-Kelvin temperatures in the framework of spin

G. Taran · W. Wernsdorfer ()


Physikalisches Institute, KIT, Karlsruhe, Germany
e-mail: gheorghe.taran@kit.edu; wolfgang.wernsdorfer@kit.edu
E. Bonet
Néel Institute, CNRS, Grenoble, France
e-mail: edgar.bonet@neel.cnrs.fr

© Springer Nature Switzerland AG 2021 979


J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_18
980 G. Taran et al.

Hamiltonian models. The application of the general formalism to mononuclear


and polynuclear complexes allows us to illustrate the power of the spin models
in explaining both static properties (e.g., magnetic bistability) and dynamic ones
(e.g., quantum tunneling of magnetization). We show how SMMs were used as a
vehicle to explore quantum phenomenologies like nonadiabatic spin transitions,
spin parity effect, the Berry phase interference, and quantum coherence while
covering milestone results that brought the field closer to providing basic com-
ponents of quantum devices. The last section is devoted to recent achievements
in the field of molecular spintronics with emphasis on basic experimental designs
that allowed the implementation of Grover’s quantum algorithms at the single-
molecule level. The successful transposition of the properties of the molecular
magnets into functional devices is a proof of the deep understanding acquired in
the two decades of scientific effort since the birth of this research field.

List of Abbreviations and Symbols

SMM Single-molecule magnets


QTM Quantum tunneling of magnetization
ZFS Zero-field splitting
SQUID Superconducting quantum interference device
LZS Landau–Zener–Stückelberg
EPR Electron paramagnetic resonance
EM Electromagnetic
SD Spin dot
RD Read-out dot
STM Scanning tunnel microscope
CNT Carbon nanotube

Introduction

Traditionally, the word magnet evokes a material in which a long-range magnetic


order arises from local exchange interactions. The observed magnetic properties,
like remanence and hysteresis, result from a collective behavior. However, devel-
opments in the last two decades showed that isolated molecules can bear large
magnetic moments that exhibit bistability like traditional magnets (Fig. 1). They
have therefore been called single-molecule magnets (SMMs).
The field of SMMs started with the discovery of the large magnetic moment
of the Mn12 O12 (CH3 COO)16 (H2 O)4 (here denoted as Mn12 -ac, Fig. 2) molecular
cluster [1] and the observation of its magnetic bistability [2]. Later, quantum
tunneling of magnetization (QTM) [3, 4] was evidenced in Mn12 -ac, as well as
ground state quantum tunneling [5] and topological quantum phase interference
effects [6] in [Fe8 O2 (OH)12 (tacn)6 ]8+ (or simply Fe8 , Fig. 2). These first discovered
19 Single-Molecule Magnets and Molecular Quantum Spintronics 981

Fig. 1 Size scale spanning atomic to nanoscale dimensions. On the far right is shown a high-
resolution transmission electron micrography of a typical 3 nm diameter cobalt nanoparticle
containing about 1000 Co atoms. The Mn84 molecule is a 4.2 nm diameter particle. Also shown for
comparison are the indicated smaller Mn nanomagnets which are drawn to scale. An alternative
means of comparison is the Néel vector (N) which is the scale shown. The arrows indicate the
magnitude of the Néel vectors for the indicated SMMs which are 7.5, 22, 61, and 168 for Mn4 ,
Mn12 , Mn30 , and Mn84 , respectively. (Reprinted from [10])

Fig. 2 Left: Structure of the Mn12 SMM with the formula [Mn12 O12 (CH3 COO)16 (H2 O)4 ] ·
2CH3 COOH·4H2 O. Middle: Structure of the Fe8 SMM with the formula [Fe8 O2 (OH)12 (tacn)6 ]8+
where tacn is a macrocyclic ligand. Right: Mononuclear terbium complex TbPc2 . The upper Pc
ligand is a mirror image of the lower Pc ligand but rotated by 45◦

SMMs lead to breakthrough observations and to this day remain among the most
investigated systems [7].
SMMs are constituted by an inner core of magnetic ions which is surrounded
by a shell of organic ligands. They come in a variety of shapes and sizes and
permit selective substitution of the ligands in order to alter the coupling to the
environment. It is also possible to replace the magnetic ions, thus changing the
magnetic properties without greatly modifying the structure and the interaction with
its surroundings.
982 G. Taran et al.

The technological interest in the field of SMMs is fueled by the desire to use
molecular structures as memory units and quantum bits (qubits). There are currently
two different approaches that try to satisfy the observed miniaturization tendencies
in the current state-of-the-art technology. The first one is a top-down approach in
which the nanometer-sized objects are obtained by reducing the dimensions of a
bulk material (a common way for obtaining magnetic nanoparticles). The second
method is a bottom-up approach [8], which for the field of SMMs means enhancing
the magnetic moment of the molecule by ion substitution or through adding
new magnetic centers to the molecule. Figure 1 shows a dimensional comparison
between the above presented methods.
One big advantage of SMMs over magnetic nanoparticles is their monodisperse
properties, as chemical synthesis yields a large number of molecules with identical
characteristics. In many cases, the SMMs can be made to form insulating monocrys-
tals; thus, the environment of each molecule is very similar. This gives access to the
properties of a single molecule from the measurements performed on the ensemble.
The found alignment between the magnetic principal axis of the molecule and the
crystallographic axes [9] is especially important when determining the intrinsic
characteristics of the magnetic centers.
SMMs belong to the mesoscopic realm (we refer to the vector space associated
with the spin degree of freedom) where properties from both the quantum and the
classical worlds transpire. The interest of working in this dimensional range comes
from the desire of using quantum properties characterizing microscopic systems
while addressing and manipulating them with simple means.
We start by presenting the spin Hamiltonian formalism, alongside introducing the
concepts essential for understanding the magnetism of molecular magnets. Then,
we show how SMMs were used as a vehicle to explore quantum phenomenologies
like quantum tunneling of magnetization (Sect. “Quantum Tunneling of Magneti-
zation”), spin parity effect and the Berry phase interference (Sect. “Spin Parity and
Quantum Phase Interference”), and quantum coherence (Sect. “Quantum Coherence
in Molecular Magnets”). The last section is devoted to recent achievements in the
field of molecular spintronics.

Spin Hamiltonian

The first step in describing the static and dynamic properties of SMMs is writing the
characteristic Hamiltonian. The application of the general formalism to mononu-
clear and polynuclear complexes will allow us to highlight both the strengths and
the range of validity of the different spin models.

Single Ion Spin Hamiltonian

A molecular complex with a single magnetic center is characterized by a total


angular momentum (J ) which results from the coupling between the spin (S) and
orbital angular momentum (L).
19 Single-Molecule Magnets and Molecular Quantum Spintronics 983

The coupling between J and the external magnetic field (H ) is modeled through
the Zeeman term:

HZ = μB μ0 H ĝJ

where ĝ is a tensor that links the total angular momentum to the magnetic moment
and μB is the Bohr magneton.
The effect of the ligand field can be accounted for through the use of the
equivalent Stevens operators [11, 12]. Thus, the total Hamiltonian is:


2S 
q
H= Bqk Oqk + HZ (1)
q=2 k=0

where Oqk are equivalent Stevens operators and Bqk are parameters related to the
ligand field and electronic structure of the ion. The Stevens operators are functions
of Jz , J− , J+ and are listed in [11, 12]. The index q represents the order of the
operators. The sum includes only even terms as only even terms generate nonzero
matrix elements. The subscript k denotes the operator’s rotational symmetry.

Transition Metal Ions


As a first example, we consider a transition metal ion compound characterized by a
weak coupling between the spin and orbital angular momentum. As a consequence,
the Hamiltonian, in a first approximation, can be written using only spin degrees of
freedom. The spin–orbit interaction, treated as a perturbation, is nonetheless very
important because it couples the magnetic ions to the surrounding organic ligands,
leading to critical contributions to the anisotropy of the molecule. By considering
only the quadratic terms from (1) in zero external magnetic field, the Hamiltonian
of the system is given by:

H = DSz2 + E(Sx2 − Sy2 ) (2)

where D and E are commonly used notations that stand for the Stevens coefficients
3B20 and B22 , respectively. D is negative when z-axis is chosen to coincide with the
easy axis of the molecule, which will be assumed in the following, and |E|  |D|
in most cases.
Considering only the first term in the above Hamiltonian, we see that the states
pertaining to the S multiplet and labeled by the quantum number m = −S...S,
are only doubly degenerate: E(±m) = −|D|m2 . For this reason, the magnetic
anisotropy constant D is often denoted as zero-field splitting (ZFS).
The discrete energy levels follow a parabola, illustrated in Fig. 3. U = |D|S 2
is called the anisotropy barrier. The two sides of the barrier correspond to opposite
orientations of the magnetic moment; thus, the ZFS term gives the approximate
energy barrier the spin has to overcome to flip its orientation.
984 G. Taran et al.

Fig. 3 Left: Energy levels of a spin S = 10 subject to the ZFS term. Right: A zoom of the energy
level diagram near an avoided level crossing where m and m are the quantum numbers of the
energy levels. PLZ is the Landau–Zener tunnel probability when sweeping the applied field from
the left to the right over the anticrossing. The greater the gap Δm,m and the slower the sweeping
rate, the higher is the tunnel probability PLZ (equation (9))

The height of the barrier and the characteristic time of the experiment lead to
the definition of the blocking temperature (TB ), as the temperature under which
the phonons do not have sufficient energy to promote reversal processes (e.g., for
Mn12 -ac, TB ∼ 4 K for a characteristic time of the experiment of 1 s). Increasing the
blocking temperature remains a central research goal in the field of SMMs [13, 14,
15] with the current record anisotropy barrier being 1760 K and TB ∼ 60 K [16,17].
The last term in (2) breaks the axial symmetry of the system and strongly affects
the low-temperature relaxation of the magnetization (Sect. “Quantum Tunneling of
Magnetization”).

Lanthanide(III) Ions
As a second example, let’s look at the lanthanide (III) ion complexes whose
magnetic properties are mainly determined by the strongly anisotropic and partially
filled 4f orbitals. Opposite to the case of transition metal ions, the angular orbital
momentum of the rare earth atoms is not quenched by the ligand field, and the spin–
orbit coupling is much stronger due to their larger nuclear charge. Thence, neither
L or S are good quantum numbers, the magnetic moment being described by a total
angular momentum, J = |L − S| . . . (L + S), where L + S is the ground state for
more than half-shell filling (Hund’s rule).
For instance, the Tb3+ ion exhibits the [Xe]4f 8 electronic structure, which leads
to a spin S = 3 and an orbital angular momentum L = 3. The ground state of the
Tb3+ ion is thus J = L+S = 6. The strong spin–orbit coupling leads to a separation
of about 2900 K between the ground state and the excited state (J = 5). We can
therefore restrict the following discussion to the ground state multiplet, J = 6,
comprised of 2J + 1 (degenerate) substates |J, mJ .
The Tb3+ ion can be embedded between two parallel phthalocyaninato (Pc)
ligand planes. The crystal field interaction generates a quantization axis oriented
perpendicular to the Pc planes. The Tb3+ ion is coordinated by four nitrogen atoms
from each ligand plane, and the upper Pc ligand (blue in Fig. 2) is a mirror reflection
of the lower one (with respect to the x-y plane) rotated by 45◦ around the z-axis. The
19 Single-Molecule Magnets and Molecular Quantum Spintronics 985

Tb3+ is therefore exposed to an antiprismatic ligand field that was modeled by the
following Hamiltonian [18]:

Hlf = B20 O20 + B40 O40 + B44 O44 + B60 O60 + B64 O64 (3)

where Bqk coefficients are written as a product of a constant depending on the


electronic shell of the ion [11] and a Akq factor characterizing the ligand field
interaction. The ligand field parameters, Akq , were determined experimentally by
fitting NMR and magnetic susceptibility measurements and can be found in [18].
The diagonalization of Hlf in the |J, mJ  eigenbasis reveals that the ligand field
partially lifts the degeneracy of the 2J + 1 substate in the ground state multiplet
J = 6. The ground state doublet, mJ = ±6, is separated from the first excited
doublet, mJ = ±5, by approximately 600 K (see [18] and Fig. 4). Then, each mJ -
doublet splits in an external magnetic field, the resulting Zeeman diagram being
depicted in Fig. 4 [19]. Therefore, at cryogenic temperatures, the TbPc2 single-
molecule magnet behaves with a good approximation as an Ising spin system.
The Tb3+ ion also carries a nuclear spin with I = 3/2 and a natural abundance
of 100%. The electric and magnetic interactions between the nuclear spin and the
electronic shell of the Tb3+ ion are represented by a hyperfine and a quadrupolar
term [20]:

HTbPc2 = Hlf + μB μ0 H ĝJ + Ahyp I · J + IP̂quad I (4)

where Ahyp = 26.7 mK is the hyperfine constant and P̂quad is the quadrupole tensor.
The hyperfine interaction, Ahyp I · J , splits the electronic states from the ground
state doublet mJ = ±6 into (2I + 1) hyperfine states, labeled by the additional
nuclear quantum number mI = −3/2 . . . 3/2.

Fig. 4 Left: Ligand field splitting of the ground state multiplet in different mononuclear lanthanide
complexes LnPc2 . (Modified from [18]). Right: Calculated Zeeman diagram for the ground state
multiplet, J = 6, in a TbPc2 SMM. (Extracted from [19])
986 G. Taran et al.

Because I > 1/2, the 159 Tb nucleus has a quadrupolar moment that couples to
the electric field gradient through IP̂quad I with the dominant term being the axial
component with magnitude Pquad = 17 mK. The tensorial nature of P̂quad reflects
the non-axial character (with respect to the easy axis of the ligand field) of the
quadrupolar interaction which together with transverse terms in Hlf was shown to be
responsible for the observed tunneling dynamics of TbPc2 in diluted single crystals
[20] (Fig. 5).

1
a)

0.5
Δ
M/M s

0
M/M s
-0.5

-1

b)
0.6
Energy (K)

0.3

-0.3

-0.04 -0.02 0.00 0.02 0.04


Bz (T)

Fig. 5 (a) Zoom on the magnetic hysteresis loop of an assembly of TbPc2 SMMs in a diluted
single crystal (1% TbPc2 in an YPc2 matrix) measured with the microSQUID setup at T =
40 mK. The insets show: (left) a zoom of a level anticrossing between two hyperfine states and
(right) the entire hysteresis loop. The red markers (right inset) indicate the continuous change
of magnetization at high fields attributed to phonon-assisted relaxation. (b) Zeeman diagram of a
TbPc2 SMM. The characteristic steps in the magnetization curve around zero field correspond to
QTM events in the single crystal and are highlighted by dashed lines. (Changed from [21])
19 Single-Molecule Magnets and Molecular Quantum Spintronics 987

Multi-spin Hamiltonian

Although a number of molecules having a single magnetic ion have been synthe-
sized [22], most SMMs contain multiple magnetic centers within their organic shell,
and their description thus requires a multi-spin Hamiltonian. For a pair of spins
coupled through exchange interaction, the Hamiltonian can be written as [23]:

Hexc = S1 Jˆ12 S2 (5)

In general, the interaction tensor Jˆ12 is not symmetric, so it’s customary to put the
above equation in a form that highlights an isotropic term (J12 S1 · S2 ) that favors
a parallel alignment of the spins, an anisotropic, symmetric component (S1 Dˆ12 S2 )
that tends to align the spins along a certain direction in space, and Dzyaloshinskii–
Moriya term (d12 · (S1 × S2 )) that works toward orienting the spins perpendicular
to each other. Usually, the isotropic term is significantly larger than the others, and
its sign will dictate the ferromagnetic or antiferromagnetic character of the ground
state.
The generalization to a system containing an arbitrary number of spins is done by
putting together the single ion anisotropies with the pairwise exchange interactions,
leading to the following multi-spin Hamiltonian:
 
H= Si D̂i Si + ˆ Sj
Si Ji,j (6)
i ij

where we considered only the quadratic term in the ligand field Hamiltonian. Even
with this simplification, the number of parameters involved can be quite large,
especially for clusters with high nuclearity. This difficulty cannot be completely
eliminated, even when one considers a very symmetric molecule. As an example,
let’s consider the archetypal Mn12 -ac molecule (Fig. 2). It contains an inner ring
of four Mn(IV) ions (S4 = 3/2) and an outer ring of eight Mn(III) (S3 = 2)
coupled together through superexchange interactions involving oxygen bridges. The
number of states that characterizes the system is (2S4 + 1)4 (2S3 + 1)8 , which yields
an overwhelming 108 eigenstates. Because the Mn12 -ac molecules crystallize in a
body-centered tetragonal crystal, the unit cell contains two equivalent Mn ions so
one independent zero-field splitting tensor is needed. When choosing a coupling
scheme with only isotropic interactions [2], another four exchange parameters are
required. Adjusting this model to experimental results, such as electron paramag-
netic resonance (EPR) measurements, is far from trivial. Nevertheless, this model is
valuable for low nuclearity systems [24]. A common way to shrink the dimension
of the Hilbert space in which we have to study the system is done by replacing pairs
of strongly interacting spins with a single equivalent spin.
988 G. Taran et al.

Giant Spin Hamiltonian

Applying the above procedure recursively to all spin pairs leads to the giant spin
approximation, which describes in an effective way the ground state multiplet. This
simplification works especially well when the temperature is low enough so that the
dominant energy in the system is the exchange interaction. In this case, the system
is always found in its spin ground state (e.g., S = 10 for Fe8 and Mn12 -ac), allowing
the use of the single ion model presented at the beginning of this section. Therefore,
we write the giant spin model:

HGS = DSz2 + gμB μ0 Hz Sz + HT (7)

where D and g are now effective parameters and HT describes non-axial interac-
tions and can be written as a function of rising and lowering spin operators (S+ ,
S− ). In order to relate the effective parameters of the giant spin model and the
parameters characterizing each individual magnetic center, one can use projection
techniques [25], where single ion anisotropies are projected onto the vector space
corresponding to the molecular ground state. This method is less accurate when the
Hilbert space is very large, like in the case of Mn12 -ac, but the model retains its use
because of its simplicity and intuitive form.
The energy range in which the giant spin model is valid can be found in the
multi-spin Hamiltonian. Indeed, the weakest exchange link gives with a good
approximation the upper temperature limit to which the predictions are expected to
be accurate. Experimentally one can observe the crossover to higher spin subspace
by determining the magnetic moment associated with the molecular cluster [26,27].

Quantum Tunneling of Magnetization

For a ligand field Hamiltonian described only by the ZFS term, the magnetic field
lifts the twofold degeneracy of the ±m energy levels. However, two different levels

(denoted by m and m ), align for specific field values Hrm,m , applied along the
magnetic principal axis:

 |D|(m + m )
Hrm,m = (8)
gμB μ0

Due to transverse terms of the spin Hamiltonian, the degeneracy can be trans-
formed into an avoided level crossing; see Fig. 3. The energy separation is called
the tunnel splitting, Δm,m , and is a central parameter that characterizes the quantum
tunneling dynamics of spins.
There are several theoretical tools that can be used to determine the tunnel
splitting, including path integral formalism [28, 29], perturbation theory [30, 31],
and numerical methods. The latter allows the possibility to consider arbitrarily
19 Single-Molecule Magnets and Molecular Quantum Spintronics 989

complex spin Hamiltonians but leaves little intuition on the system’s behavior. Thus,
a combination of the above techniques is to be preferred when performing both
quantitative and qualitative analyses of the magnetic properties of SMMs.
At an avoided level crossing, the eigenvectors of the Hamiltonian are a linear
combination of the base vectors that correspond to negative and positive spin
projections. This means that in these states, the magnetization of the spin has a
nonzero probability to be found on either side of the barrier. This behavior is
called spin tunneling, that is, the spin is in quantum resonance between the opposite
orientation states.
At a constant field tuned at an avoided level crossing, a spin initially situated
on one side of the well will oscillate coherently between the mixed states with a

characteristic angular frequency, ωTm,m , related to the tunnel splitting through the

relation: Δm,m = 2h̄ωTm,m . The field interval where this behavior is predicted to
Δ 
happen is given by [23]: δH0 = gμB μ0m,m
|m−m | . This quantity is called the bare tunnel
−9
width and can be as small as 10 T. However, environmental interactions broaden
the observed width of the resonance and hinder coherent oscillations [32,33,34,35].

Landau–Zener–Stückelberg (LZS) Model

The nonadiabatic transition between the states of a two-level system was first
discussed by Landau, Zener, and Stückelberg [36, 37, 38]. The original work by
Zener concentrates on the electronic states of a bi-atomic molecule, while Landau
and Stückelberg considered two atoms that undergo a scattering process. Their
solution to the time-dependent Schrödinger equation of a two-level system driven
through resonance has been applied to many physical systems, and it became an
important tool for studying tunneling transitions. The LZS model was used to
analyze spin tunneling in nanoparticles and clusters [39, 40, 41, 42]. The tunneling
probability, PLZ , between the states m and m , after sweeping the applied field at a
constant rate, α, through the resonance (Fig. 3), is given by:
 
δH0
PLZ = 1 − exp −π ωT (9)
α

With the LZS model in mind, we can now start to understand qualitatively the
hysteresis loops of SMMs (Fig. 6), which exhibit steps of fast relaxation separated
by regions where the magnetization is almost constant. The steps happen at specific
fields where the levels from both sides of the well are mixed by transverse terms in
the spin Hamiltonian, as discussed in the previous section.
Let us start at a large negative magnetic field Hz . At very low temperatures,
all molecules are in the m = 10 ground state (Fig. 6). As the applied field Hz is
ramped down to zero, all molecules will stay in the m = 10 ground state. When
ramping the field over the Δ10,−10 region, at Hz ≈ 0, there is a Landau–Zener
probability, P10,−10 , for the spins to tunnel from the m = 10 to the m = −10
990 G. Taran et al.

Fig. 6 Top: Hysteresis loops


of a single crystal of Fe8
molecular clusters at 0.04 K
and different field sweep
rates. The loops display a
series of steps, separated by
plateaus. Arrows indicate the
resonance transitions
highlighted in the figure
below. Bottom: Zeeman
diagram of the S = 10
manifold of Fe8 as a function
of the field applied along the
easy axis. From bottom to
top, the levels are labeled
with quantum numbers
m = ±10, ±9, . . .. The levels
cross at equidistant field
values given by
μ0 Hz ≈ n × 0.22 T with
n = 1, 2, 3 . . .. The arrows
are explained in the text

state. P10,−10 depends on the sweeping rate (9); the slower the sweeping rate, the
larger is the tunneling probability. This is clearly demonstrated in the hysteresis loop
measurements showing larger steps for slower sweeping rates [43,6]. When the field
Hz is now further increased, there is a remaining fraction of molecules in the now
metastable m = 10 state. The next chance to escape from this state is when the field
reaches the Δ10,−9 region. There is a Landau–Zener tunnel probability P10,−9 to
tunnel from the m = 10 to the m = −9 state. As m = −9 is an excited state, the
molecules in this state relax quickly to the m = −10 state by emitting a phonon. An
analogous procedure happens when the applied field reaches the Δ10,−10+n regions
(n = 2, 3, . . . ) until all molecules are in the m = −10 ground state, that is, the
magnetic moment of all molecules is reversed. As phonon emission can only change
the molecule state by Δm = ±1 or ±2, there is a phonon cascade for higher applied
fields.
Figure 5 shows the magnetic hysteresis loop, measured by microSQUID tech-
nique [44], of a crystal containing TbPc2 SMMs randomly distributed in a
diamagnetic, isostructural matrix formed by YPc2 molecules, with [TbPc2 ]/[YPc2 ]
ratio of 1%. Upon sweeping the magnetic field from −1 T up to positive fields as
small as 0.05 T, approximately 75% of the TbPc2 SMMs undergo quantum tunneling
19 Single-Molecule Magnets and Molecular Quantum Spintronics 991

transitions, resulting in sharp steps in the magnetization curve. The QTM transitions
take place between mixed states of nuclear and electronic origin; thus, both spin
projections can change [20]. The remaining SMM reverse their magnetic moment
at larger magnetic fields by a direct relaxation process [45].
When using the above formalism for a quantitative analysis, one should keep in
mind that the LZS model is exact only for an isolated spin. Deviations from the ideal
coherent dynamics are due to environmental interactions of both elastic (dephasing)
and inelastic (relaxation and excitation) nature [21]. Thus, in order to try to search an
agreement with LZS theory, one should use large sweeping rates, so that in the time
the resonance is swept, the local environmental field does not change significantly.
The model also doesn’t include relaxation through phonons so one should work at
very low temperatures and focus on ground state tunneling.
LZS theory was successfully used to determine the tunnel splitting in molecular
systems such as Fe8 [6] and Mn4 [43]. As mentioned, the agreement with the
experiment must be searched in the fast sweeping rate regime, which has the
drawback of a small sensitivity. To overcome this difficulty, the resonance can be
swept repeatedly. This way, the probability for the spin to remain in the original
state, for small variations of the magnetization, after n back-and-forth sweeps, is
proportional to (1 − 2nPLZ ), with PLZ given by (9).

Spin Parity and Quantum Phase Interference

When discussing quantum tunneling of magnetization, it was emphasized that


transverse terms remove the degeneracy of the eigenstates at a level crossing
(Hz = Hr ) and thus promote relaxation. Actually, the general problem of eigenvalue
degeneracy is discussed by the von Neumann–Wigner theorem [46,47] which states
the need of at least two parameters (e.g., the components of the magnetic field), to
control the degeneracy property of a Hermitian matrix. In this section, we discuss
the cases when the degeneracy is predicted theoretically and in some cases observed
experimentally through the absence of tunneling at level intersections.

Spin Parity
The general form of a transverse term of order n is Bn S±n ; thus, a transition between
 n
the levels m and m is made possible by applying S± operator an integer number of
times. Consequently, the degeneracy is removed only when the change in the spin
projection is a multiple of the perturbation’s order: n|(m − m ).
The above quantum tunneling selection rule is called spin parity effect and has
its origin in the symmetry of the non-axial Hamiltonian term. For example, in the
case of the Mn3 cluster [48], with a C3 rotational symmetry, level splitting should
occur only if the selection rule m − m = 3k is satisfied.
Another interesting example is the observation of Kramer’s degeneracy in half-
integer spin molecular clusters [49], which tells us that in zero applied magnetic
field, the ground state is at least doubly degenerate. This degeneracy can be lifted
992 G. Taran et al.

by a transverse field (a first-order perturbation) of external or internal origin (e.g.,


dipolar and hyperfine field).
In general, the selection rules are rarely observed. Possible explanations involve
asymmetries introduced by crystal defects, solvent disorder [50], abovementioned
dipolar and hyperfine interactions, or Dzyaloshinskii–Moriya interaction [51]
fuelled by low molecular symmetry and strong exchange coupling. In the case when
one gets to control the above factors, the spin parity effect can be clearly evidenced,
as seen for Mn3 [48].

Quantum Phase Interference


Suppression of tunneling can also occur for certain values of the transverse field,
without involving the above-presented selection rule.
The interference effect in spin systems was predicted theoretically [28] and then
observed experimentally [6] using LZS approach and is generally called the Berry
phase interference. The mechanics behind the explanation of this phenomena makes
it similar to the observation of critical current oscillations in the Josephson junctions
[52, 53] and Bohm–Aharonov oscillations [54].
A semiclassical description, which has a rather intuitive picture associated with
it, describes the tunneling motion of the spin with the help of the path integral
formalism [28]. The initial and final states are represented by two points on
the Bloch sphere (Fig. 7). The interference is due to a nonzero phase difference

Fig. 7 Unit sphere showing degenerate minima A and B which are joined by two tunnel paths
(heavy lines). The hard, medium, and easy axes are taken in the x-, y-, and z-direction, respectively.
The constant transverse field Htrans for tunnel splitting measurements is applied in the xy-plane
at an azimuth angle ϕ. At zero applied field H = 0, the giant spin reversal results from the
interference of two quantum spin paths of opposite directions in the easy anisotropy yz-plane.
For transverse fields along an axis perpendicular to the easy axis, by using Stokes’ theorem, it has
been shown that the path integrals can be converted in an area integral, yielding that destructive
interference – that is, a quench of the tunneling rate – occurs whenever the area enclosed by the
two paths is kπ/S, where k is an odd integer. (Reprinted from [6])
19 Single-Molecule Magnets and Molecular Quantum Spintronics 993

Fig. 8 Measured tunnel splitting Δ as a function of transverse field for (a) several azimuth angles
ϕ at m = ±10 and (b) ϕ ≈ 0◦ , as well as for quantum transition between m = −10 and (10 − n).
Note the parity effect that is analogous to the suppression of tunneling predicted for half-integer
spins. It should also be mentioned that internal dipolar and hyperfine fields hinder a complete
quench of Δ which is predicted for an isolated spin. (Reprinted from [6])

acquired by moving in the parametric space along a closed path containing these
two points.
The magnetic field has the ability to modulate the phase by modifying the paths,
so an interference pattern is observed as the transverse field is increased.
Figure 8 depicts tunnel splitting oscillations of a Fe8 complex showing equally
spaced minima as a function of the transverse field. The difference between the
maxima and the minima of the tunnel splitting can be as large as one order of
magnitude.
The anisotropy of the Fe8 molecule, similar to many low symmetry molecules, is
modeled by adding a biaxial crystal field (HT = E(Sx2 − Sy2 )) to the uniaxial term.
Therefore, when the tunnel splitting between the levels S and S − n is measured,
the spin parity effect is observed (the second-order transverse anisotropy term
forbids odd resonances at zero field). Also, a monotonic increase of Δ with n is
observed: the lower is the energy barrier, the higher are the tunneling rates.
The predicted field separation between two consecutive minima of the tunnel
splitting, when we consider only the biaxial term, is given by:

2 
ΔH = 2E(E + D) (10)
gμB

Using the anisotropy parameters determined in [55], one obtains a field separa-
tion of 0.26 T, which is smaller than the experimental value of 0.4 T. In order to
account for this difference, higher order terms must be considered. This is easily
done in the numerical approach that involves the diagonalization of the system’s
Hamiltonian.
Other systems in which it was possible to observe the Berry phase interference
include Mn12 complexes [56,57,58], Mn4 [59], and dimer molecular magnets [60].
994 G. Taran et al.

Quantum Coherence in Molecular Magnets

The study of coherence in molecular magnets is very important for the potential
applications in the field of quantum information processing. Here we briefly account
on the quest for the observation of coherent time evolution of molecular spins under
electromagnetic (EM) radiation.

Resonant Photon Absorption

In EPR experiments involving molecular magnets [55, 61, 62], transitions between
the m and m states of the S multiplet happen when the energy of the incoming
photons is equal to the energy difference between the states: hν = |E(m) − E(m )|,
obeying the selection rule Δm = ±1. An important question that had to be answered
concerns the effect of the photon-induced excitations on tunneling.
The first important results [63] were obtained when a Fe8 complex was
investigated using micro-Hall bars in a dilution refrigerator. The study proved that
the lifetime of the excited states is large enough (relative to the tunneling time)
for an enhanced relaxation to be observed. In order to clearly show the photon-
assisted quantum tunneling regime, circular polarized microwave radiation was used
because it allows to select the Δm = +1 or Δm = −1 transitions. Thus, an
asymmetric hysteresis loop is observed. From the dependence of the transition rates
on the power of the microwave source used, it was shown that the effective spin
temperature (the parameter which describes the occupation number of the excited
states) depends linearly on the power of the EM field. Subsequent experimental
work further proved the photon-assisted tunneling regime [64,65,66] leading toward
eventual observation of Rabi oscillations [62].

Rabi Oscillations

The coherent evolution of the system between two eigenstates coupled by the EM
field is described by the Rabi model [68,69]. If the system’s initial state is |m, then
the probability to find it in the state |m , at time t, is proportional to sin2 (ΩR t).
The Rabi frequency, ΩR , is proportional to the amplitude of the perturbative field.
This property is used in the spin-echo measurement protocols [70] to determine the
characteristic relaxation times of the system.
The longitudinal relaxation time (T1 ) is obtained from the recovery of the
equilibrium magnetization after an inversion pulse, π − T − π2 − τ − π − τ − echo,
is applied, where the variable is T and τ is being kept constant. T1 is directly
connected to the coupling of the spins to the phonon bath and thus can be
significantly long at mK temperatures.
The second important characteristic time is the phase coherence time (T2 ) and is
determined by spin–spin interactions. As the name suggests, it tells us the time over
which the memory of the quantum phase is preserved, so the quantum properties of
the spin can be exploited. Employing a similar Hahn-echo sequence, π2 − τ − π −
19 Single-Molecule Magnets and Molecular Quantum Spintronics 995

Fig. 9 Comparison of
state-of-the-art T2 values of
notable molecular and
solid-state electronic spin
qubits. (Extracted from [67])

τ − echo, with τ variable, T2 is obtained from the echo signals that are fitted to a
stretched exponential, I (τ ) = I (0) exp(−(2τ/T2 )x ).
The interaction with environmental spins of nuclear and electronic origin,
phonons, and photons represents the main sources of decoherence [71] in SMMs
[72, 73] and thus limits T2 . In order to reduce the dipolar interactions, one usually
chooses a system with a small collective spin [65, 74, 62] because the magnitude
of the interactions scales as the square of the spin magnitude. Then, SMMs can be
diluted without greatly effecting their individual properties [74, 62, 75]. Also, the
initial choice of a molecular complex with S = 1/2 [74] avoided the problems
associated with the distribution of the anisotropy axis.
The above measurements led to the first observations of a long phase coherence
time in Cr7 Ni (S = 1/2) and Cr7 Mn (S = 1) clusters [74]. Afterward, Rabi
oscillations were observed in V15 [62, 76], and T2 on μs scale was measured in
Fe8 [77] through the application of a high transverse magnetic field.
Recent developments showed that a strict control over lattice rigidity and
hyperfine interactions can lead to significantly large phase coherence times, even
when compared to other qubit candidates ( [75] and Fig. 9). The coherence also
has been shown to be preserved at room temperature [78, 79]. These are valuable
observations that encourage possible spintronics applications.

Molecular Quantum Spintronics

Molecular quantum spintronics is a relatively new research field that combines


spintronics, molecular electronics, and quantum computing with the aim of creating
new spintronics molecular devices that exploit the quantum properties seen at the
microscopic level [80,81,82]. These devices facilitate the read-out and manipulation
of the spin states pertaining to a molecular magnet, leading to structures that can
perform basic quantum operations.
The idea of using SMMs as magnetic centers in spintronics devices is supported
by their unique characteristics, namely, weak spin–orbit coupling in transition metal
996 G. Taran et al.

ion compounds, tunable environmental interactions that can lead to a long coher-
ence time (Sect. “Quantum Coherence in Molecular Magnets”), and chemically
controlled functionalities like switchability with light or electric field [83]. The
coupling to external structures can be facilitated by choosing an appropriate ligand.
Delocalized bonds, which mediate the interaction between the magnetic ions, often
imply great conduction properties.
The above-outlined properties allowed the realization of some essential circuit
elements, like [81] molecular spin transistors [84, 85], molecular spin valves [86]
and spin filters, and molecular double-dot devices.

Direct Coupling

The first configuration that we consider involves a direct coupling mechanism in


which the molecules are connected to metallic electrodes through chemical bonds
(Fig. 10). This facilitates a strong coupling between the conduction electrons and
the molecular spin but also implies a strong back action on the spin state.
The setup was tested experimentally when a low-temperature scanning tunnel
microscope (STM) was used to probe molecular magnets deposited on a conducting
surface, leading to the observation of reversible chiral switching [87] and electrical

Fig. 10 Schematic
representation of different
device geometries for
molecular spintronics
devices. (a) Three-terminal
spin-dot device. This is a
direct coupling scheme, in
which the current flows
through the spin dot (SD). (b)
Three-terminal double-dot
device. This is an indirect
coupling scheme, in which
the current flows through a
second nonmagnetic quantum
dot, the read-out dot (RD),
which is coupled with the
spin dot (SD) via exchange
interaction (J). (c)
Supramolecular spin valve
device, with two SD’s
coupled to the RD. The
current in the RD is therefore
sensitive to relative spin
orientation in the two SD’s.
(d) Multiterminal multi-dot
device. (Reprinted from [45])
19 Single-Molecule Magnets and Molecular Quantum Spintronics 997

current control over the spin state [88]. The spin-polarized STM allows direct space
resolved observations of spin-split orbitals [89].
Another configuration is represented by a three terminal device where a molecu-
lar magnet bridges the gap of a break junction. A gate electrode is used to tune the
transport properties of the conduction electrons [90]. First implementation of this
design involved a Mn12 compound bound covalently to gold electrodes. To explore
different conduction regimes, weak binding ligands were also used. In both cases
the peripheral group acted as a tunnel barrier, helping to conserve the magnetic
properties of the cluster.
By knowing the magnetic properties of the redox species that are formed
when the electrons pass through the molecule, one can compare the information
obtained from spectroscopic transport measurements and established experimen-
tal methods like: magnetization measurements, EPR or neutron scattering. First
studies involving Mn12 in a magnetic field showed that it exhibits transistor like
properties [90, 91]. Both degeneracy in zero field and non-linear behavior of the
excitations as a function of field are typical of tunneling via a magnetic molecule.
However, follow up experiments indicated an alteration of the Mn12 magnetic
properties during the deposition on gold electrodes [92]. In contrast to Mn12 , both
Fe4 [93] and TbPc2 [94] are known to preserve their magnetic properties upon
deposition.
In the following we show how such a three terminal device, containing TbPc2
deposited on electromigrated gold electrodes, was used to read-out not only the
electronic spin state but also the nuclear spin state of Tb3+ [94].

Read-Out of a Single Nuclear Spin


The Tb3+ center was shown to have a stable redox state which suggests that
the current is not flowing through the ion. Instead, it is more likely that the Pc
ligands, that contain a conjugated π system, are the conducting medium. Transport
measurements revealed a single charge degeneracy with a weak spin-1/2 Kondo
effect. When studying the position of the Kondo peak as a function of the applied
field (supplemental information of [94]), an exchange interaction of about 0.35 T
between a quantum dot with S = 1/2 and the Tb3+ ion was found. This relatively
strong interaction indicates that the quantum dot and the Tb3+ ion are fairly close
together, suggesting a binding geometry between the molecule and the electrodes as
shown in Fig. 11. The spin states of the Tb3+ ion are not influenced by the charge
state of the ligands, thus indicating that the Pc ligands behave as read-out dots. The
response to a varying magnetic field was studied, using the charge degeneracy point
as the working point.
The experiment showed single jumps of the conductance as the field was swept
over the zero-field region. These sudden changes in conductance are connected
to the relaxation of the spin through quantum tunneling. In order to construct
the statistics of the magnetization reversal, the resonance region was swept
22,000 times. The resulting histogram, showed in Fig. 12, illustrates that the
magnetization reversal occurs at four distinct field values. The results are in perfect
agreement with the discussion on quantum tunneling behavior of TbPc2 presented in
998 G. Taran et al.

Fig. 11 Artist view of (left) a nuclear spin qubit transistor based on a single TbPc2 molecular
magnet, coupled to source, drain, and gate (not shown) electrodes, and (right) a three-spin-dot
device. (Reprinted from [85])

Fig. 12 Zeeman diagram


and nuclear spin detection
procedure. Top: Zeeman
diagram of the TbPc2
molecular magnet, showing
the hyperfine split electronic
spin ground state doublet as a
function of the external
magnetic field H|| parallel to
the easy-axis of
magnetization. The ligand
field-induced avoided level
crossings (colored rectangles)
allow for tunneling of the
electron spin. Middle: The
jumps of the conductance of
the read-out quantum dot
during magnetic field sweeps
are nuclear spin dependent.
Bottom: Histogram of the
positions of about 75,000
conductance jumps, showing
four non-overlapping
Gaussian-like peaks. Each
conductance jump can be
assigned to a specific nuclear
spin state

Sect. “Landau–Zener–Stückelberg (LZS) Model”. Thus, the position of a particular


transition tells us not only the state of the electronic spin but also the nuclear spin
state.
Because there is no overlap between the distribution of the switching fields, it is
possible to determine the state of the nuclear spin through a single measurement.
19 Single-Molecule Magnets and Molecular Quantum Spintronics 999

The lifetime of the nuclear states can be determined by doing subsequent mea-
surements separated by a waiting time [85]. A T1 of order of tens of seconds was
observed, which shows that this is a non-invasive method.
The hyperfine Stark effect was used to manipulate the nuclear spin within a
single molecule with an individual relaxation time (T1 ) exceeding 20 s [85]. Rabi,
Ramsey and Hahn echo experiments (Fig. 13), performed on the 159 Tb nuclear
spin [95], compare well with similar experiments involving P or Bi defects in
Si. Ramsey T2 -times are 300 microseconds with π/2 pulse-lengths of 100 ns, i.e.,
about 1000 spin manipulations are possible before decoherence sets in. This allowed
the implementation of Grover’s algorithm using the four states of the nuclear
spin [96].

Indirect Coupling

In order to avoid the strong back action characterizing the above presented devices,
the molecule can be coupled to the electrodes by indirect means, that is, by establish-
ing a link between the SMM (a spin dot) and a non-magnetic molecular conductor
(read-out dot) connected to the electrodes (Fig. 10). The coupling between the spin
dot and read-out dot is usually realized through an exchange interaction and can be
modulated by a gate voltage.
Among possible candidates for read-out dots (e.g., nanowires, carbon nanotubes,
nano-SQUIDs and ligands), carbon nanotubes are special due to their outstanding
structural, mechanical and electrical properties. Because a carbon nanotube is
essentially a 1D molecular conductor, with a diameter of the same order as a
molecular magnet, a good coupling is easily achieved. At low temperature, it
presents electronic properties like Coulomb blockade [97] and Kondo effect [98].
Thus, carbon nanotube conductance is sensitive to any charge fluctuations including
the spin reversal processes.
A number of devices with a carbon nanotube as the detector and TbPc2 as the
spin dot showed that the indirect coupling mechanism allows the determination of
both the electronic and nuclear spin of Tb3+ ion. Urdampilleta et al. [86] showed
that a carbon nanotube functionalized with two SMMs has a supramolecular valve
behavior (Fig. 14). Indeed, the two TbPc2 molecules behave as a spin polarizer-
analyzer system exhibiting a strong magnetoresistive effect.
Mediated by exchange interaction, the magnetic moment of each molecule
induces a localized spin-polarized dot in the carbon nanotube quantum dot that can
be controlled by a magnetic field (Fig. 14). Depending on the relative orientation of
the molecular spins, a high- or low-conductance regime can be observed. A butterfly
hysteresis loop, characteristic of spin valve devices, with a magnetoresistance up to
300% has been measured for temperatures smaller than 1 K. For a more detailed
description, the reader may refer to Urdampilleta et al. [86].
Increasing the number of SMMs coupled to the carbon nanotube (Fig. 10)
and addressing each individual molecule through a local gate may allow the
implementation of complex quantum computation protocols.
1000

Fig. 13 Qubit coherent manipulations. Each color is a transition between consecutive levels: red being the transition between the ground state and the first
excited level, and the next two in green and blue. (a) Rabi oscillations, the frequency of these oscillations can be tuned from 1.5 to 9 MHz. The maximum
visibility is respectively 0.6, 0.9 and 0.6. (b) By recording the maximum of the visibility of the Rabi oscillations as function of the detuning, the shape of the
three transitions can be measured. (c) Ramsey fringes give a coherence time of the order of 0.3 ms for the three transitions. (d) The Results of the Hahn-echo
measurement show that we are far from the limit T1 = 2T2 . This indicates that the decoherence mechanism is not due to the relaxation process. (Changed
from [96])
G. Taran et al.
19 Single-Molecule Magnets and Molecular Quantum Spintronics 1001

Fig. 14 Spin valve behavior in a supramolecular spintronics device based on a carbon nanotube
quantum dot functionalized with TbPc2 SMMs. Figures from [86,99]: (a) Butterfly hysteresis loop
at T = 40 mK. (b) Antiparallel spin configuration: the spin state in dot A is reversed with respect
to that of dot B. The energy mismatch between levels with different spins results in a current
blockade. (c) Parallel spin configuration for both molecules A and B. Energy levels with same spin
are aligned allowing electron transport through the carbon nanotube

Quantum Algorithms

The research aimed at using molecular magnets for qubit encoding started mainly
after the theoretical proposal of Leuenberger and Loss [100, 101], followed by
others [102,103,104,105,106,107]. We shall follow DiVincenzo [108] to summarize
the steps taken toward successful implementation of a quantum computer using
molecular magnets.
To use molecular magnets as qubits, external constrains (e.g., temperature,
electric or magnetic field) should be applied in order to confine the system to a
subset of two levels (|0 and |1). The scalability of the system is facilitated by
the synthetic bottom-up fabrication process that guarantees cheap production of
identical molecular units.
The implementation of proposed quantum algorithms [109, 110, 111, 112]
involves the use of one-qubit and two-qubit gates. The former represents a rotation
on the Bloch sphere. The output of the gate, after realizing an electric or magnetic
coupling to the spin and applying an EM pulse sequence (Sect. “Quantum Coher-
ence in Molecular Magnets”), is the state: α|0 + β|1. A strong-coupling regime, at
high temperatures, between a molecular spin ensemble and microwave resonators,
has been achieved. The possibility of coupling strongly with single molecules has
been put forward, and experiments are in progress. This opens a way to develop
scalable architectures using molecular spins coupled to quantum circuits [113, 114,
115, 116, 117]. Moreover, by molecular engineering of the crystal field, molecular
spins can also be manipulated by electric fields [118, 119, 120, 84, 121].
A two-qubit gate can be implemented by controlling the exchange interaction
between the spins. Schemes and compounds for switchable effective qubit–qubit
interactions in the presence of permanent exchange couplings are now available
1002 G. Taran et al.

[122, 123]. Spin entanglement between and within molecules was shown by
different techniques. Molecular spin clusters/arrays offer an incredible variety of
spin topologies to test entanglement at the molecular scale [119].
It is necessary that the system remains in a coherent state (preserve α and
β), for a time considerably larger than the computational clock time. This is an
important requirement for the system to be amenable to quantum error corrections.
As noted in Sect. “Quantum Coherence in Molecular Magnets”, long coherence time
(T2 ) at low temperature has been reported for many ensembles of molecular spin
systems with T2 approaching 1 ms in nuclear spin-free environments [74, 124, 72].
Recent reports have shown microsecond coherence times and Rabi oscillations at
room temperature [78, 125, 75, 126, 127, 79]. Strategies to protect spin states from
decoherence (e.g., via atomic clock transitions) have been experimentally tested by
fine engineering of molecular states and levels [128]. The development of theoretical
schemes to implement quantum error correction codes in molecules with multiple
spin degrees of freedom has also been started [129, 130, 131, 132, 133].
Another important requirement is the possibility to initialize and read out the
qubit state. As seen in previous sections, the read-out of a single molecular
spin located at a tunnel junction or on a CNT/graphene quantum dot has been
demonstrated [134, 86, 135].
All these achievements are important milestones in molecular magnetism. They
bring this research field closer to being able to provide the future basic components
of quantum devices [136].

Conclusion

This chapter gives an overview of the main phenomenologies related to magnetic


properties of molecular magnets and presents some important achievements in the
field of molecular spintronics, aimed toward the implementation of a quantum com-
puter. We highlight the symbiotic relationship between fundamental research and
technological application as a central feature of this research field, the successful
transposition of the properties of the molecular magnets into functional devices
being a proof of the deep understanding acquired in the last decades.

References
1. Caneschi, A., Gatteschi, D., Sessoli, R., Barra, A.L., Brunel, L.C., Guillot, M.: Alternating
current susceptibility, high field magnetization, and millimeter band epr evidence for a ground
s = 10 state in [Mn12 O12 (CH3 COO)16 (H2 O)4 ].2CH3 COOH.4H2 O. J. Am. Chem. Soc.
113(15), 5873–5874 (1991)
2. Sessoli, R., Tsai, H.-L., Schake, A.R., Sheyi, W., Vincent, J.B., Folting, K., Gatteschi, D.,
Christou, G., Hendrickson, D.N.: High-spin molecules:[mn12o12 (o2cr) 16 (h2o) 4]. J. Am.
Chem. Soc. 115(5), 1804–1816 (1993)
3. Friedman, J.R., Sarachik, M., Tejada, J., Ziolo, R.: Macroscopic measurement of resonant
magnetization tunneling in high-spin molecules. Phys. Rev. Lett. 76(20), 3830 (1996)
19 Single-Molecule Magnets and Molecular Quantum Spintronics 1003

4. Thomas, L., Lionti, F., Ballou, R., Gatteschi, D., et al.: Macroscopic quantum tunnelling of
magnetization in a single crystal of nanomagnets. Nature 383(6596), 145 (1996)
5. Sangregorio, C., Ohm, T., Paulsen, C., Sessoli, R., Gatteschi, D.: Quantum tunneling of the
magnetization in an iron cluster nanomagnet. Phys. Rev. Lett. 78(24), 4645 (1997)
6. Wernsdorfer, W., Sessoli, R.: Quantum phase interference and parity effects in magnetic
molecular clusters. Science 284(5411), 133–135 (1999)
7. Friedman, J.R., Sarachik, M.P.: Single-molecule nanomagnets. Annu. Rev. Condens. Matter
Phys. 1(1), 109–128 (2010)
8. Lu, W., Lieber, C.M.: Nanoelectronics from the bottom up. Nature Mater. 6(11), 841–850
(2007)
9. Wernsdorfer, W., Chakov, N., Christou, G.: Determination of the magnetic anisotropy axes of
single-molecule magnets. Phys. Rev. B 70(13), 132413 (2004)
10. Tasiopoulos, A.J., Vinslava, A., Wernsdorfer, W., Abboud, K.A., Christou, G.: Giant single-
molecule magnets: a {Mn84} torus and its supramolecular nanotubes. Angewandte Chemie
116(16), 2169–2173 (2004)
11. Stevens, K.: Matrix elements and operator equivalents connected with the magnetic properties
of rare earth ions. Proc. Phys. Soc. Sect. A 65(3), 209 (1952)
12. Abragam, A., Bleaney, B.: Electron paramagnetic resonance of transition ions. OUP, Oxford
(2012)
13. Waldmann, O.: A criterion for the anisotropy barrier in single-molecule magnets. Inorg.
Chem. 46(24), 10035–10037 (2007)
14. Ishikawa, N., Sugita, M., Ishikawa, T., Koshihara, S.-Y., Kaizu, Y.: Mononuclear lanthanide
complexes with a long magnetization relaxation time at high temperatures: a new category of
magnets at the single-molecular level. J. Phys. Chem. B 108(31), 11265–11271 (2004)
15. Langley, S.K., Wielechowski, D.P., Moubaraki, B., Murray, K.S.: Enhancing the magnetic
blocking temperature and magnetic coercivity of {Cr III2 Ln III2} single-molecule magnets
via bridging ligand modification. Chem. Commun. 52(73), 10976–10979 (2016)
16. Goodwin, C.A., Ortu, F., Reta, D., Chilton, N.F., Mills, D.P.: Molecular magnetic hysteresis
at 60 kelvin in dysprosocenium. Nature 548(7668), 439 (2017)
17. Layfield, R., Guo, F.-S., Day, B., Chen, Y.-C., Tong, M.-L., Mansikamäkki, A.: A dysprosium
metallocene single-molecule magnet functioning at the axial limit. Angew. Chem. Int. Ed.
56(38), 11445–11449 (2017)
18. Ishikawa, N., Sugita, M., Okubo, T., Tanaka, N., Iino, T., Kaizu, Y.: Determination of
ligand-field parameters and f-electronic structures of double-decker bis (phthalocyaninato)
lanthanide complexes. Inorg. Chem. 42(7), 2440–2446 (2003)
19. Ishikawa, N., Sugita, M., Wernsdorfer, W.: Quantum tunneling of magnetization in lanthanide
single-molecule magnets: bis (phthalocyaninato) terbium and bis (phthalocyaninato) dyspro-
sium anions. Angew. Chem. Int. Ed. 44(19), 2931–2935 (2005)
20. Taran, G., Bonet, E., Wernsdorfer, W.: The role of the quadrupolar interaction in the tunneling
dynamics of lanthanide molecular magnets. J. Appl. Phys. 125(14), 142903 (2019)
21. Taran, G., Bonet, E., Wernsdorfer, W.: Decoherence measurements in crystals of molecular
magnets. Phys. Rev. B 99(18), 180408 (2019)
22. Zheng, Y.-Z., Zheng, Z., Chen, X.-M.: A symbol approach for classification of molecule-
based magnetic materials exemplified by coordination polymers of metal carboxylates. Coord.
Chem. Rev. 258, 1–15 (2014)
23. Gatteschi, D., Sessoli, R., Villain, J.: Molecular Nanomagnets, vol. 5. Oxford University Press
on Demand, Oxford (2006)
24. Hill, S., Datta, S., Liu, J., Inglis, R., Milios, C.J., Feng, P.L., Henderson, J.J., del Barco,
E., Brechin, E.K., Hendrickson, D.N.: Magnetic quantum tunneling: insights from simple
molecule-based magnets. Dalton Trans. 39(20), 4693–4707 (2010)
25. Gatteschi, D., Bencini, A.: Electron Paramagnetic Resonance of Exchange Coupled Systems.
Springer, Berlin/Heidelberg (1990)
26. Tupitsyn, I., Barbara, B.: Quantum tunneling of magnetization in molecular complexes with
large spins–effect of the environment. Magnetism: molecules to materials: 5 Volumes Set,
pp. 109–168 (2001)
1004 G. Taran et al.

27. Barbara, B.: Quantum tunneling of the collective spins of single-molecule magnets: from
early studies to quantum coherence. In: Molecular Magnets, pp. 17–60. Springer, Berlin/Hei-
delberg (2014)
28. Garg, A.: Topologically quenched tunnel splitting in spin systems without kramers’ degener-
acy. EPL (Europhys. Lett.) 22(3), 205 (1993)
29. Schilling, R.: Quantum spin-tunneling: a path integral approach. In: Quantum Tunneling of
Magnetization: QTM’94, pp. 59–76. Springer, Dordrecht (1995)
30. Garanin, D.: Spin tunnelling: a perturbative approach. J. Phys. A: Math. Gen. 24(2), L61
(1991)
31. Hartmann-Boutron, F., Politi, P., Villain, J.: Tunneling and magnetic relaxation in mesoscopic
molecules. Int. J. Modern Phys. B 10(21), 2577–2637 (1996)
32. Wernsdorfer, W., Ohm, T., Sangregorio, C., Sessoli, R., Mailly, D., Paulsen, C.: Observation
of the distribution of molecular spin states by resonant quantum tunneling of the magnetiza-
tion. Phys. Rev. Lett. 82(19), 3903 (1999)
33. Wernsdorfer, W., Caneschi, A., Sessoli, R., Gatteschi, D., Cornia, A., Villar, V., Paulsen, C.:
Effects of nuclear spins on the quantum relaxation of the magnetization for the molecular
nanomagnet fe 8. Phys. Rev. Lett. 84(13), 2965 (2000)
34. Alonso, J.J., Fernández, J.F.: Tunnel window’s imprint on dipolar field distributions. Phys.
Rev. Lett. 87(9), 097205 (2001)
35. Tupitsyn, I., Stamp, P., Prokof’ev, N.: Hole digging in ensembles of tunneling molecular
magnets. Phys. Rev. B 69(13), 132406 (2004)
36. Landau, L.: Zur theorie der energieubertragung. II. Phys. Z. Sowjetunion 2(46), 1–13 (1932)
37. Zener, C.: Non-adiabatic crossing of energy levels. In: Proceedings of the Royal Society of
London A: Mathematical, Physical and Engineering Sciences, vol. 137, pp. 696–702. The
Royal Society (1932)
38. Stueckelberg, E.: Theory of inelastic collisions between atoms(theory of inelastic collisions
between atoms, using two simultaneous differential equations). Helv. Phys. Acta (Basel) 5,
369–422 (1932)
39. Miyashita, S.: Dynamics of the magnetization with an inversion of the magnetic field. J. Phys.
Soc. Jpn. 64(9), 3207–3214 (1995)
40. Miyashita, S.: Observation of the energy gap due to the quantum Tunneling making use of the
landau-zener mechanism. J. Phys. Soc. Jpn. 65(8), 2734–2735 (1996)
41. Thorwart, M., Grifoni, M., Hänggi, P.: Tunneling and vibrational relaxation in driven multi-
level systems. Phys. Rev. Lett. 85, quant-ph/9912024, 860 (2000)
42. Leuenberger, M.N., Loss, D.: Incoherent zener tunneling and its application to molecular
magnets. Phys. Rev. B 61(18), 12200 (2000)
43. Wernsdorfer, W., Bhaduri, S., Vinslava, A., Christou, G.: Landau-zener tunneling in the
presence of weak intermolecular interactions in a crystal of mn 4 single-molecule magnets.
Phys. Rev. B 72(21), 214429 (2005)
44. Wernsdorfer, W.: From micro-to nano-squids: applications to nanomagnetism. Superconduc-
tor Sci. Technol. 22(6), 064013 (2009)
45. Ganzhorn, M., Wernsdorfer, W.: Molecular quantum spintronics using single-molecule
magnets. In: Molecular Magnets, pp. 319–364. Springer, Berlin/Heidelberg (2014)
46. Bruno, P.: Berry phase, topology, and degeneracies in quantum nanomagnets. Phys. Rev. Lett.
96(11), 117208 (2006)
47. Demkov, Y.N., Kurasov, P.B.: Von neumann-wigner theorem: Level repulsion and degenerate
eigenvalues. Theor. Math. Phys. 153(1), 1407–1422 (2007)
48. Henderson, J., Koo, C., Feng, P., Del Barco, E., Hill, S., Tupitsyn, I., Stamp, P., Hendrickson,
D.: Manifestation of spin selection rules on the quantum tunneling of magnetization in a
single-molecule magnet. Phys. Rev. Lett. 103(1), 017202 (2009)
49. Wernsdorfer, W., Bhaduri, S., Boskovic, C., Christou, G., Hendrickson, D.: Spin-parity
dependent tunneling of magnetization in single-molecule magnets. Phys. Rev. B 65(18),
180403 (2002)
50. Del Barco, E., Kent, A.D., Hill, S., North, J., Dalal, N., Rumberger, E., Hendrickson, D.,
Chakov, N., Christou, G.: Magnetic quantum tunneling in the single-molecule magnet mn12-
acetate,” J. Low Temp. Phys. 140(1–2), 119–174 (2005)
19 Single-Molecule Magnets and Molecular Quantum Spintronics 1005

51. De Raedt, H., Miyashita, S., Michielsen, K., Machida, M.: Dzyaloshinskii-moriya interactions
and adiabatic magnetization dynamics in molecular magnets. Phys. Rev. B 70(6), 064401
(2004)
52. Josephson, B.D.: Possible new effects in superconductive tunnelling. Phys. Lett. 1(7), 251–
253 (1962)
53. Koelle, D., Kleiner, R., Ludwig, F., Dantsker, E., Clarke, J.: High-transition-temperature
superconducting quantum interference devices. Rev. Modern Phys. 71(3), 631 (1999)
54. Aharonov, Y., Bohm, D.: Significance of electromagnetic potentials in the quantum theory.
Phys. Rev. 115(3), 485 (1959)
55. Barra, A.-L., Debrunner, P., Gatteschi, D., Schulz, C.E., Sessoli, R.: Superparamagnetic-
like behavior in an octanuclear iron cluster. EPL (Europhys. Lett.) 35s(2), 133
(1996)
56. Wernsdorfer, W., Soler, M., Christou, G., Hendrickson, D.: Quantum phase interference
(berry phase) in single-molecule magnets of [mn 12] 2-. J. Appl. Phys. 91(10), 7164–7166
(2002)
57. Wernsdorfer, W., Chakov, N., Christou, G.: Quantum phase interference and spin-parity in mn
12 single-molecule magnets. Phys. Rev. Lett. 95(3), 037203 (2005)
58. Adams, S., da Silva Neto, E.H., Datta, S., Ware, J., Lampropoulos, C., Christou, G.,
Myaesoedov, Y., Zeldov, E., Friedman, J.R.: Geometric-phase interference in a mn 12 single-
molecule magnet with fourfold rotational symmetry. Phys. Rev. Lett. 110(8), 087205 (2013)
59. Quddusi, H.M., Liu, J., Singh, S., Heroux, K., Del Barco, E., Hill, S., Hendrickson, D.:
Asymmetric berry-phase interference patterns in a single-molecule magnet. Phys. Rev. Lett.
106(22), 227201 (2011)
60. Ramsey, C.M., Del Barco, E., Hill, S., Shah, S.J., Beedle, C.C., Hendrickson, D.N.: Quantum
interference of tunnel trajectories between states of different spin length in a dimeric
molecular nanomagnet. Nat. Phys. 4(4), 277–281 (2008)
61. Mossin, S., Stefan, M., ter Heerdt, P., Bouwen, A., Goovaerts, E., Weihe, H.: Fourth-order
zero-field splitting parameters of [mn (cyclam) br2] br determined by single-crystal w-band
epr. Appl. Magn. Reson. 21(3–4), 587–596 (2001)
62. Bertaina, S., Gambarelli, S., Mitra, T., Tsukerblat, B., Müller, A., Barbara, B.: Quantum
oscillations in a molecular magnet. Nature 453(7192), 203–206 (2008)
63. Sorace, L., Wernsdorfer, W., Thirion, C., Barra, A.-L., Pacchioni, M., Mailly, D., Barbara,
B.: Photon-assisted tunneling in a fe 8 single-molecule magnet. Phys. Rev. B 68(22), 220407
(2003)
64. Bal, M., Friedman, J.R., Suzuki, Y., Mertes, K., Rumberger, E., Hendrickson, D., Myasoedov,
Y., Shtrikman, H., Avraham, N., Zeldov, E.: Photon-induced magnetization reversal in the fe
8 single-molecule magnet. Phys. Rev. B 70(10), 100408 (2004)
65. Wernsdorfer, W., Müller, A., Mailly, D., Barbara, B.: Resonant photon absorption in the low-
spin molecule v15. EPL (Europhys. Lett.) 66(6), 861 (2004)
66. Petukhov, K., Wernsdorfer, W., Barra, A.-L., Mosser, V.: Resonant photon absorption in fe
8 single-molecule magnets detected via magnetization measurements. Phys. Rev. B 72(5),
052401 (2005)
67. Moreno-Pineda, E., Godfrin, C., Balestro, F., Wernsdorfer, W., Ruben, M.: Molecular spin
qudits for quantum algorithms. Chem. Soc. Rev. 47(2), 501–513 (2018)
68. Rabi, I.I.: Space quantization in a gyrating magnetic field. Phys. Rev. 51(8), 652 (1937)
69. Grifoni, M., Hänggi, P.: Driven quantum tunneling. Phys. Rep. 304(5), 229–354 (1998)
70. Schweiger, A., Jeschke, G.: Principles of Pulse Electron Paramagnetic Resonance. Oxford
University Press on Demand, Oxford (2001)
71. Zurek, W.H.: Decoherence, einselection, and the quantum origins of the classical. Rev.
Modern Phys. 75(3), 715 (2003)
72. Takahashi, S., Tupitsyn, I., Van Tol, J., Beedle, C., Hendrickson, D., Stamp, P.: Decoherence
in crystals of quantum molecular magnets. Nature 476(7358), 76–79 (2011)
73. Shim, J., Bertaina, S., Gambarelli, S., Mitra, T., Müller, A., Baibekov, E., Malkin, B.,
Tsukerblat, B., Barbara, B.: Decoherence window and electron-nuclear cross relaxation in
the molecular magnet v 15. Phys. Rev. Lett. 109(5), 050401 (2012)
1006 G. Taran et al.

74. Ardavan, A., Rival, O., Morton, J.J., Blundell, S.J., Tyryshkin, A.M., Timco, G.A., Winpenny,
R.E.: Will spin-relaxation times in molecular magnets permit quantum information process-
ing? Phys. Rev. Lett. 98(5), 057201 (2007)
75. Zadrozny, J.M., Niklas, J., Poluektov, O.G., Freedman, D.E.: Millisecond coherence time in
a tunable molecular electronic spin qubit. ACS Cent. Sci. 1(9), 488–492 (2015)
76. Yang, J., Wang, Y., Wang, Z., Rong, X., Duan, C.-K., Su, J.-H., Du, J.: Observing quantum
oscillation of ground states in single molecular magnet. Phys. Rev. Lett. 108(23), 230501
(2012)
77. Takahashi, S., van Tol, J., Beedle, C.C., Hendrickson, D.N., Brunel, L.-C., Sherwin, M.S.:
Coherent manipulation and decoherence of s = 10 single-molecule magnets. Phys. Rev. Lett.
102(8), 087603 (2009)
78. Bader, K., Dengler, D., Lenz, S., Endeward, B., Jiang, S.-D., Neugebauer, P., van Slageren, J.:
Room temperature quantum coherence in a potential molecular qubit. Nat. Commun. 5, 5304
(2014)
79. Atzori, M., Tesi, L., Morra, E., Chiesa, M., Sorace, L., Sessoli, R.: Room-temperature
quantum coherence and rabi oscillations in vanadyl phthalocyanine: toward multifunctional
molecular spin qubits. J. Am. Chem. Soc. 138(7), 2154–2157 (2016)
80. Cleuziou, J.-P., Wernsdorfer, W., Bouchiat, V., Ondarçuhu, T., Monthioux, M.: Carbon
nanotube superconducting quantum interference device. Nat. Nanotechnol. 1(1), 53–59
(2006)
81. Bogani, L., Wernsdorfer, W.: Molecular spintronics using single-molecule magnets. Nat.
Mater. 7(3), 179–186 (2008)
82. Roch, N., Florens, S., Bouchiat, V., Wernsdorfer, W., Balestro, F.: Quantum phase transition
in a single-molecule quantum dot. Nature 453(7195), 633–637 (2008)
83. Sanvito, S., Rocha, A.R.: Molecular-spintronics: the art of driving spin through molecules. J.
Comput. Theor. Nanosci. 3(5), 624–642 (2006)
84. Zyazin, A.S., van den Berg, J.W., Osorio, E.A., van der Zant, H.S., Konstantinidis, N.P.,
Leijnse, M., Wegewijs, M.R., May, F., Hofstetter, W., Danieli, C., et al.: Electric field
controlled magnetic anisotropy in a single molecule. Nano Lett. 10(9), 3307–3311 (2010)
85. Thiele, S., Vincent, R., Holzmann, M., Klyatskaya, S., Ruben, M., Balestro, F., Wernsdorfer,
W.: Electrical readout of individual nuclear spin trajectories in a single-molecule magnet spin
transistor. Phys. Rev. Lett. 111(3), 037203 (2013)
86. Urdampilleta, M., Klyatskaya, S., Cleuziou, J.-P., Ruben, M., Wernsdorfer, W.: Supramolec-
ular spin valves. Nat. Mater. 10(7), 502–506 (2011)
87. Fu, Y.-S., Schwöbel, J., Hla, S.-W., Dilullo, A., Hoffmann, G., Klyatskaya, S., Ruben, M.,
Wiesendanger, R.: Reversible chiral switching of bis (phthalocyaninato) terbium (iii) on a
metal surface. Nano Lett. 12(8), 3931–3935 (2012)
88. Komeda, T., Isshiki, H., Liu, J., Zhang, Y.-F., Lorente, N., Katoh, K., Breedlove, B.K.,
Yamashita, M.: Observation and electric current control of a local spin in a single-molecule
magnet. Nat. Commun. 2, 217 (2011)
89. Schwöbel, J., Fu, Y., Brede, J., Dilullo, A., Hoffmann, G., Klyatskaya, S., Ruben, M.,
Wiesendanger, R.: Real-space observation of spin-split molecular orbitals of adsorbed single-
molecule magnets. Nat. Commun. 3, 953 (2012)
90. Heersche, H., De Groot, Z., Folk, J., Van Der Zant, H., Romeike, C., Wegewijs, M., Zobbi,
L., Barreca, D., Tondello, E., Cornia, A.: Electron transport through single mn 12 molecular
magnets. Phys. Rev. Lett. 96(20), 206801 (2006)
91. Jo, M.-H., Grose, J.E., Baheti, K., Deshmukh, M.M., Sokol, J.J., Rumberger, E.M., Hen-
drickson, D.N., Long, J.R., Park, H., Ralph, D.: Signatures of molecular magnetism in
single-molecule transport spectroscopy. Nano Lett. 6(9), 2014–2020 (2006)
92. Mannini, M., Sainctavit, P., Sessoli, R., Cartier dit Moulin, C., Pineider, F., Arrio, M.-A.,
Cornia, A., Gatteschi, D.: Xas and xmcd investigation of mn12 monolayers on gold. Chem.–
A Eur. J. 14(25), 7530–7535 (2008)
93. Burzurí, E., Zyazin, A., Cornia, A., Van der Zant, H.: Direct observation of magnetic
anisotropy in an individual fe 4 single-molecule magnet. Phys. Rev. Lett. 109(14), 147203
(2012)
19 Single-Molecule Magnets and Molecular Quantum Spintronics 1007

94. Vincent, R., Klyatskaya, S., Ruben, M., Wernsdorfer, W., Balestro, F.: Electronic read-out of
a single nuclear spin using a molecular spin transistor. Nature 488(7411), 357–360 (2012)
95. Thiele, S., Balestro, F., Ballou, R., Klyatskaya, S., Ruben, M., Wernsdorfer, W.: Electrically
driven nuclear spin resonance in single-molecule magnets. Science 344(6188), 1135–1138
(2014)
96. Godfrin, C., Ferhat, A., Ballou, R., Klyatskaya, S., Ruben, M., Wernsdorfer, W., Balestro, F.:
Operating quantum states in single magnetic molecules: implementation of grover’s quantum
algorithm. Phys. Rev. Lett. 119(18), 187702 (2017)
97. Charlier, J.-C., Blase, X., Roche, S.: Electronic and transport properties of nanotubes. Rev.
Modern Phys. 79(2), 677 (2007)
98. Nygård, J., Cobden, D.H., Lindelof, P.E.: Kondo physics in carbon nanotubes. Nature
408(6810), 342–346 (2000)
99. Urdampilleta, M., Klyatskaya, S., Ruben, M., Wernsdorfer, W.: Landau-zener tunneling of a
single tb 3+ magnetic moment allowing the electronic read-out of a nuclear spin. Phys. Rev.
B 87(19), 195412 (2013)
100. Leuenberger, M.N., Loss, D.: Quantum computing in molecular magnets. Nature 410(6830),
789–793 (2001)
101. Leuenberger, M.N., Loss, D.: Grover algorithm for large nuclear spins in semiconductors.
Phys. Rev. B 68(16), 165317 (2003)
102. Meier, F., Levy, J., Loss, D.: Quantum computing with spin cluster qubits. Phys. Rev. Lett.
90(4), 047901 (2003)
103. Troiani, F., Ghirri, A., Affronte, M., Carretta, S., Santini, P., Amoretti, G., Piligkos, S.,
Timco, G., Winpenny, R.: Molecular engineering of antiferromagnetic rings for quantum
computation. Phys. Rev. Lett. 94(20), 207208 (2005)
104. Carretta, S., Santini, P., Amoretti, G., Troiani, F., Affronte, M.: Spin triangles as optimal units
for molecule-based quantum gates. Phys. Rev. B 76(2), 024408 (2007)
105. Troiani, F., Affronte, M.: Molecular spins for quantum information technologies. Chem. Soc.
Rev. 40(6), 3119–3129 (2011)
106. Lehmann, J., Gaita-Arino, A., Coronado, E., Loss, D.: Spin qubits with electrically gated
polyoxometalate molecules. Nat. Nanotechnol. 2(5), 312–317 (2007)
107. Bartolomé, J., Luis, F., Fernández, J.F.: Molecular magnets. Phys. Appl. Springer (2014)
108. DiVincenzo, D.P., et al.: The physical implementation of quantum computation. arXiv
preprint quant-ph/0002077 (2000)
109. Deutsch, D., Jozsa, R.: Rapid solution of problems by quantum computation. In: Proceedings
of the Royal Society of London A: Mathematical, Physical and Engineering Sciences,
vol. 439, pp. 553–558. The Royal Society (1992)
110. Steane, A.: Quantum computing. Rep. Progress Phys. 61(2), 117 (1998)
111. Grover, L.K.: A fast quantum mechanical algorithm for database search. In: Proceedings of
the Twenty-Eighth Annual ACM Symposium on Theory of Computing, pp. 212–219. ACM
(1996)
112. Shor, P.W.: Polynomial-time algorithms for prime factorization and discrete logarithms on a
quantum computer. SIAM Rev. 41(2), 303–332 (1999)
113. Abe, E., Wu, H., Ardavan, A., Morton, J.J.: Electron spin ensemble strongly coupled to a
three-dimensional microwave cavity. Appl. Phys. Lett. 98(25), 251108 (2011)
114. Chiorescu, I., Groll, N., Bertaina, S., Mori, T., Miyashita, S.: Magnetic strong coupling in a
spin-photon system and transition to classical regime. Phys. Rev. B 82(2), 024413 (2010)
115. Clauss, C., Bothner, D., Koelle, D., Kleiner, R., Bogani, L., Scheffler, M., Dressel, M.:
Broadband electron spin resonance from 500 mhz to 40 ghz using superconducting coplanar
waveguides. Appl. Phys. Lett. 102(16), 162601 (2013)
116. Jenkins, M., Hümmer, T., Martínez-Pérez, M.J., García-Ripoll, J., Zueco, D., Luis, F.:
Coupling single-molecule magnets to quantum circuits. New J. Phys. 15(9), 095007 (2013)
117. Carretta, S., Chiesa, A., Troiani, F., Gerace, D., Amoretti, G., Santini, P.: Quantum informa-
tion processing with hybrid spin-photon qubit encoding. Phys. Rev. Lett. 111(11), 110501
(2013)
1008 G. Taran et al.

118. Trif, M., Troiani, F., Stepanenko, D., Loss, D.: Spin electric effects in molecular antiferro-
magnets. Phys. Rev. B 82(4), 045429 (2010)
119. Troiani, F., Bellini, V., Candini, A., Lorusso, G., Affronte, M.: Spin entanglement in
supramolecular structures. Nanotechnology 21(27), 274009 (2010)
120. Nossa, J., Islam, M., Canali, C.M., Pederson, M.: First-principles studies of spin-orbit and
dzyaloshinskii-moriya interactions in the {Cu 3} single-molecule magnet. Phys. Rev. B 85(8),
085427 (2012)
121. Graham, M.J., Zadrozny, J.M., Shiddiq, M., Anderson, J.S., Fataftah, M.S., Hill, S.,
Freedman, D.E.: Influence of electronic spin and spin–orbit coupling on decoherence in
mononuclear transition metal complexes. J. Am. Chem. Soc. 136(21), 7623–7626 (2014)
122. Timco, G.A., Carretta, S., Troiani, F., Tuna, F., Pritchard, R.J., Muryn, C.A., McInnes, E.J.,
Ghirri, A., Candini, A., Santini, P., et al.: Engineering the coupling between molecular spin
qubits by coordination chemistry. Nat. Nanotechnol. 4(3), 173–178 (2009)
123. Ferrando-Soria, J., Pineda, E.M., Chiesa, A., Fernandez, A., Magee, S.A., Carretta, S.,
Santini, P., Vitorica-Yrezabal, I.J., Tuna, F., Timco, G.A., et al.: A modular design of
molecular qubits to implement universal quantum gates. Nat. Commun. 7, 11377 (2016)
124. Ardavan, A., Bowen, A.M., Fernandez, A., Fielding, A.J., Kaminski, D., Moro, F., Muryn,
C.A., Wise, M.D., Ruggi, A., McInnes, E.J., et al.: Engineering coherent interactions in
molecular nanomagnet dimers. arXiv preprint arXiv:1510.01694 (2015)
125. Zadrozny, J.M., Niklas, J., Poluektov, O.G., Freedman, D.E.: Multiple quantum coherences
from hyperfine transitions in a vanadium (iv) complex. J. Am. Chem. Soc. 136(45), 15841–
15844 (2014)
126. Tesi, L., Lucaccini, E., Cimatti, I., Perfetti, M., Mannini, M., Atzori, M., Morra, E., Chiesa,
M., Caneschi, A., Sorace, L., et al.: Quantum coherence in a processable vanadyl complex:
new tools for the search of molecular spin qubits. Chem. Sci. 7(3), 2074–2083 (2016)
127. Gómez-Coca, S., Urtizberea, A., Cremades, E., Alonso, P.J., Camón, A., Ruiz, E., Luis,
F.: Origin of slow magnetic relaxation in kramers ions with non-uniaxial anisotropy. Nat.
Commun. 5 4300 (2014)
128. Shiddiq, M., Komijani, D., Duan, Y., Gaita-Ariño, A., Coronado, E., Hill, S.: Enhancing
coherence in molecular spin qubits via atomic clock transitions. Nature 531(7594), 348–351
(2016)
129. Hodges, J.S., Yang, J.C., Ramanathan, C., Cory, D.G.: Universal control of nuclear spins via
anisotropic hyperfine interactions. Phys. Rev. A 78(1), 010303 (2008)
130. Santini, P., Carretta, S., Troiani, F., Amoretti, G.: Molecular nanomagnets as quantum
simulators. Phys. Rev. Lett. 107(23), 230502 (2011)
131. Zhang, Y., Ryan, C.A., Laflamme, R., Baugh, J.: Coherent control of two nuclear spins using
the anisotropic hyperfine interaction. Phys. Rev. Lett. 107(17), 170503 (2011)
132. Whitehead, G.F., Cross, B., Carthy, L., Milway, V.A., Rath, H., Fernandez, A., Heath, S.L.,
Muryn, C.A., Pritchard, R.G., Teat, S.J., et al.: Rings and threads as linkers in metal–organic
frameworks and poly-rotaxanes. Chem. Commun. 49(65), 7195–7197 (2013)
133. Ueda, A., Suzuki, S., Yoshida, K., Fukui, K., Sato, K., Takui, T., Nakasuji, K., Morita, Y.:
Hexamethoxyphenalenyl as a possible quantum spin simulator: an electronically stabilized
neutral π radical with novel quantum coherence owing to extremely high nuclear spin
degeneracy. Angew. Chem. Int. Ed. 52(18), 4795–4799 (2013)
134. Candini, A., Klyatskaya, S., Ruben, M., Wernsdorfer, W., Affronte, M.: Graphene spintronic
devices with molecular nanomagnets. Nano Lett. 11(7), 2634–2639 (2011)
135. Urdampilleta, M., Klayatskaya, S., Ruben, M., Wernsdorfer, W.: Magnetic interaction
between a radical spin and a single-molecule magnet in a molecular spin-valve. ACS Nano
9(4), 4458–4464 (2015)
136. Ghirri, A., Troiani, F., Affronte, M.: Quantum computation with molecular nanomagnets:
achievements, challenges, and new trends. In: Molecular Nanomagnets and Related Phenom-
ena, pp. 383–430. Springer (2014)
19 Single-Molecule Magnets and Molecular Quantum Spintronics 1009

Taran Gheorghe born in R. Moldova in 1990. He finished his


bachelor and master degree in physics at University “Alexandru
Ioan Cuza,” Romania. Currently, he is pursuing a doctoral degree
in molecular magnetism at the Karlsruhe Institute of Technology
in Germany.

Edgar Bonet studied physics and computer science at the “École


Normale Supérieure” in Lyon. He completed his PhD at the Louis
Néel laboratory, in Grenoble, France, in 1999. After a 2-year
postdoctoral position at Cornell University, he came back to
Grenoble as a permanent researcher. He now works at the Institut
Néel on nanomagnetism and molecular magnetism.

Wolfgang Wernsdorfer born in Germany in 1966, studied


physics at the University of Würzburg and École Normale
Supérieure in Lyon. In 1993, he became a doctoral researcher at
the Low Temperature Laboratory and the Laboratoire de Mag-
netism in Grenoble, France, where he then stayed as a researcher.
In 2004, he became directeur de recherche at the Institut NEEL in
Grenoble. In 2016, he accepted a position as Humboldt Professor
at the Institute of Physics and the Institute of Quantum Materials
and Technologies at the KIT.
Magnetic Nanoparticles
20
Sara A. Majetich

Contents
Ideal Single-Domain Nanoparticles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1012
How Real Magnetic Nanoparticles Can Be Different . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1015
Magnetic Nanoparticle Preparation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1017
General Nanoparticle Formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1017
Liquid Phase Syntheses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1018
Applications of Magnetic Nanoparticles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1021
Important Magnetic Characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1021
Magnetic Separation and Manipulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1023
Magnetic Hyperthermia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1024
Magnetic Particle Imaging (MPI) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1026
Contrast Agents for Magnetic Resonance Imaging (MRI) . . . . . . . . . . . . . . . . . . . . . . . . . 1026
Ferrofluids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1028
Magnetic Recording Media . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1029
Nonbiomedical Topics of Recent Interest . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1030
L10 FePt Nanoparticles for Magnetic Recording Media . . . . . . . . . . . . . . . . . . . . . . . . . . . 1030
Core-Shell NPs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1032
Surface Effects and Spin Canting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1033
Frontiers and Future Directions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1035
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1037

Abstract

The physics of ideal single-domain nanoparticles is reviewed, followed by a


discussion of how the behavior of real particles can differ. The main synthetic
approaches are surveyed, and the advantages of different methods are identified.
The magnetic properties of the nanoparticles are related to their use in applica-

S. A. Majetich ()
Physics Department, Carnegie Mellon University, Pittsburgh, PA, USA
e-mail: sara@cmu.edu

© Springer Nature Switzerland AG 2021 1011


J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_20
1012 S. A. Majetich

tions including magnetic separation, ferrofluids, magnetic recording media, and


biomedicine. Topics of recent interest are discussed, along with future directions.

Monodomain magnetic nanoparticles (NPs) may be treated to first order as


macrospins, but their detailed behavior can be quite complex. They have been
extensively studied because they serve as models for magnetic recording media, and
they are used in magnetic fluids and biomedical applications. Here we will focus
on small magnetic nanoparticles prepared by chemical means that can be dispersed
in liquids and moved by external forces. We start with a short description of the
physics that unifies the field of magnetic nanoparticles and enables comparison
of different materials. Next we discuss the differences between real and idealized
NPs. Different approaches to NP synthesis are presented, along with explanations of
constraints imposed by their intended use. The roles of magnetic NPs in applications
in biomedicine, ferrofluids, and magnetic recording are discussed, along with the
potential for future applications. Finally there is a survey of some recent trends in
magnetic NP research.
There have been numerous reviews on magnetic nanoparticles, from the perspec-
tive of theoretical models [1], collective behavior [2], and biomedical applications
[3–5]. For reasons of space, some subfields will be limited or omitted, including
nanoparticle synthesis, lithographically patterned nanoparticles, and atomistic sim-
ulations. This review is intended as a guide to the magnetism that unites the field
of magnetic nanoparticles, together with a survey of their applications and current
research trends.

Ideal Single-Domain Nanoparticles

In 1930 Frenkel and Dorfman first suggested the existence of ferromagnetic domains
[6]. Bitter confirmed them by dusting fine magnetic particles over the surface of a
ferromagnetic material [7]. We now know that in bulk ferromagnetic materials, the
spins are parallel due to short-range quantum mechanical exchange interactions.
There are also long-range magnetostatic interactions, and the combination of
exchange and magnetostatic interactions leads to the spontaneous formation of
magnetic domains, where locally the spins are parallel but over the sample the
domain magnetization directions form magnetic flux closure paths to minimize the
overall energy.
Below a critical size, it is not energetically favorable to form a domain wall, and
the particle is said to be monodomain. An ellipsoidal particle of a ferromagnetic
or ferromagnetic material will be uniformly magnetized. The magnetic moment of
the particle μ = μ0 Ms V, where Ms is the saturation magnetization and V is the
particle volume. The maximum monodomain diameter can be estimated from the
parameters of the bulk ferromagnetic (FM) or ferromagnetic (FiM) material [8]:
20 Magnetic Nanoparticles 1013

Fig. 1 A monodomain
magnetic nanoparticle. Here Easy Axis
µ
the direction of the particle
moment μ depends on the
effective anisotropy, which H
determines the easy axis
direction, and the external
magnetic field H

72(Ak)1/2
dcr = (1)
μ0 Ms2

Here A is the exchange stiffness and K is the anisotropy. A and Ms depend on the
material, while the anisotropy has contributions from the crystallographic structure
and direction and particle shape. Figure 1 shows the case for spherical particle of a
uniaxial material, where the energy is minimized when the magnetization lies along
an easy axis. For more complex cases with biaxial and cubic anisotropy, see Ref. [1].
Maximum single-domain sizes for different materials based on bulk parameters
have been tabulated elsewhere [9]. While Eq. 1 is useful as a guide, there are many
reasons it can be inaccurate. Surface spins have fewer near neighbors for exchange
coupling. Symmetry breaking at the surface leads to an additional anisotropy
contribution that can distort the NP spin configuration [10]. If the oxidation state
of surface atoms differs from that of the interior, as with metal NPs, the effective
Ms will be reduced compared with that for the bulk metal, even if the spins
are not canted. For materials with very large magnetocrystalline anisotropy, the
theoretical dcr could be hundreds of nanometers, but structural defects such as grain
boundaries or antiphase boundaries limit the experimentally observed domain size.
At the opposite extreme, with very low magnetocrystalline anisotropy, there is little
energetic cost if they are not perfectly collinear. Here the domain wall width is a
better estimate of the single-domain threshold.
In bulk ferromagnets magnetization reversal occurs through domain wall motion,
but in single-domain particles, the exchange-coupled spins remain parallel to one
another as they rotate coherently, as a macrospin [11, 12]. The switching field or
coercivity, Hc , is a maximum for particles at the largest monodomain size [13] and
can be predicted from an energy barrier model [14], where the barrier to reversal is
given by

 
Ms H 2
E = KV 1 − . (2)
2K

Hc drops until it reaches zero at the superparamagnetic particle size dSP :


1014 S. A. Majetich

 1/3
6kT B
dSP = , (3)
π K ln (τmeas /τ0 )

where τ 0 is the inverse of the Larmor precession frequency, ∼10−9 s, and τ meas is
the measurement time. The threshold temperature for superparamagnetism is called
the blocking temperature, TB , defined as

TB = KV ln (τmeas /τ0 ) /k, (4)

but it too will depend on the measurement time. For a superconducting quantum
interference effect device (SQUID) magnetometer, the time is ∼100 s, but Möss-
bauer spectroscopy and neutron scattering have measurement times of ∼10−7 s or
shorter. In magnetic hyperthermia, the frequency ranges from 100 kHz to 1 MHz,
and particles that look superparamagnetic based on SQUID magnetometry may be
blocked over the shorter time scales [2].
By definition, superparamagnetic particles do not interact with each other, and
the magnetization of an ensemble of superparamagnets, relative to its value at zero
temperature, is a Langevin function:

M 1
= L(x) = coth(x) − , (5)
M(0) x

where x = μH/kT. Figure 2 shows an example of a Langevin function fit. It is


important that the sample is dilute (∼0.1 vol.%), or magnetostatic interactions of
the particles will distort the curve and will typically lead to underestimation of the
particle moment μ.

Fig. 2 Langevin function fit of a dilute sample of 8.5 nm Fe3 O4 nanoparticles. Below the blocking
temperature of ∼65 K, the two branches of the loop have been averaged. (unpublished results)
20 Magnetic Nanoparticles 1015

The dynamics of relaxation for superparamagnets are predicted by the Néel


model [12, 15], an ensemble of identical, rigid, noninteracting particles has a
magnetization that decays exponentially with time. The Néel-Brown model assumes
an Arrhenius law for the escape rate from a metastable state. Rearranging Eq. 4, the
rate for a given temperature is

τ 1 = τ0 1 exp (−KV /kB T ) . (6)

This is a reasonable approximation if the particle spends the majority of its time
in the local or global ground state, rather than in a transition state. To understand
dynamics on shorter time scales (100 ns or less), the Landau-Lifshitz-Gilbert
equation provides a more accurate picture of precessional switching [16].

How Real Magnetic Nanoparticles Can Be Different

The view of magnetic NPs as macrospins is a helpful starting point, but real particles
can be more complex in many ways. Here we describe a few of the most common
differences.
The nanoparticles need not be chemically uniform. Most nanoparticles have a
surfactant or polymer coating that forms a diffusion barrier to oxygen, but does not
prevent oxidation of sensitive materials. Sudden exposure to air can cause magnetic
metal NPs of Fe or Co to catch fire, but gradual exposure leads to a core-shell
structure. Metals oxidize by O2 adsorption and dissociation on the surface, followed
by metal ion migration to react and nucleate the metal oxide. The oxide shells
surrounding magnetic metal cores are usually polycrystalline, with ∼1–2 nm grains
having different crystallographic orientations, relative to the interior. In extreme
cases, the Kirkendall effect is observed, in which the particles become hollow, and
all that remains is a porous metal oxide shell [17]. Since Fe and Co oxides can be
magnetically ordered. Bulk Fe3 O4 and γ-Fe2 O3 are ferromagnetic, while FeO is
antiferromagnetic. Bulk CoO and Co3 O4 are antiferromagnetic. When these phases
form a thin shell around a metallic core, there can be exchange bias between metal
core and oxide shell spins [18], just as in bilayer thin films [19, 20]. However, the
behavior is much more complex due to the polycrystalline nature of the shell, which
leads to a wide variety of superexchange interactions within a single particle.
Many NPs show reduced Ms compared with that of the bulk material. In some
cases thin magnetic films and NPs are said to have a “dead layer” without ascribing
an origin, but there are many hypotheses about the underlying reasons. Some are
inherent in the symmetry breaking at the surface. This introduces a new contribution
to the anisotropy that can distort the spins from the completely parallel configuration
of the ideal monodomain particle. Negative surface anisotropy favors spins that lie
in the plane of the surface [10, 21]. Positive surface anisotropy favors perpendicular
orientation [10]; at moderate values this leads to surface distortion akin to a
point dipole pattern. If the NP cores are amorphous, the surface anisotropy will
dominate [22]. Many groups have assumed that the surface anisotropy contribution
1016 S. A. Majetich

is inversely proportional to the NP diameter, but in reality many different factors


contribute to an effective anisotropy [23].
Not all NPs show significant reduction in Ms [24–26], but in real NPs there are
two additional factors that can reduce the surface magnetization: structural disorder
and chemical variations. A spin glass model was proposed to explain the behavior
of ball-milled ferrite NPs [27, 28]. Materials that have a combination of ferro- and
antiferro-magnetic exchange interactions, such as spinel ferrites and L10 FePt [29],
are most susceptible to disorder-induced magnetization reductions. However, this
disorder could occur either at the surface or in the interior, in the form of antiphase
boundaries [25]. NP surfaces have a unique form of disorder due to their curved
surfaces; a broader range of near neighbor exchange interactions, relative to the
interior, can lead to a magnetic frustration that would reduce the net magnetization.
Most NPs have something bonded to their surface to prevent agglomeration, and
unless the species is physisorbed, it will alter the chemical composition of the
surface. An example of this would be oleic acid bonding to iron oxide NPs, where
the two oxygen ions of the carboxylic acid bind with a surface Fe ion. Surfactant
coverage varies with the preparation method, and also with the type of surface. Oleic
acid binds more strongly to (100), leading to faster growth along <111> directions
[30]. The particle shape can be varied depending on the proportion of oleic acid
and weakly binding oleyl amine [30]. Without complete passivation, there will be a
chemically and magnetically inhomogeneous surface. In recent years attention has
been paid to the inhomogeneity of oxygen distribution in iron oxide NPs. Many
groups have observed ∼5% oxygen vacancies in “magnetite-like” NPs [25, 31].
While much closer to Fe3 O4 than to γ-Fe2 O3 in composition, these particles did not
undergo a Verwey transition near 120 K that is characteristic of bulk magnetite. The
electrical conductivity of Fe3 O4 is three orders of magnitude greater than that of
γ-Fe2 O3 and is associated with charge hopping between Fe2+ and Fe3+ or polaron
formation [32]. At high temperature, γ-Fe2 O3 has the structure of Fe3 O4 with
oxygen vacancies and without Fe2+ ions. The addition of oxygen vacancies to a
Fe3 O4 NP would change the chemistry and reduce the conductivity. Magnetite NPs
that undergo a Verwey transition have been routinely prepared by aqueous methods
[33], but here particles are rougher and less monodisperse than those from inert
atmosphere methods. The latter approach has been modified using CO to generate
small, smooth NPs that have a metal-insulator transition [34], and it is hoped that
syntheses with less toxic reagents will be developed in the future.
While most magnetic nanoparticles are nominally spherical, some degree of
shape anisotropy is common. Bulk Fe3 O4 has [35] easy axis directions. However,
a 5% distortion of a 10 nm magnetite sphere is sufficient to make it uniaxial
due to shape anisotropy. Faceting is commonly observed in NPs. The rate of
growth of (111) and (100) faces depends on the type and concentration of the
surfactant. Truncated octahedra are frequently observed. While cube-shaped NPs
have been prepared for a variety of materials [36–38], the cubes generally have
rounded corners. Needle-shaped γ-Fe2 O3 nanoparticles where the shape anisotropy
makes them uniaxial, with coercivity ∼800 Oe, were used as magnetic recording
media for tapes and for the first hard disk drive (RAMAC). However, there is
20 Magnetic Nanoparticles 1017

less research focus on elongated (acicular) magnetic nanoparticles today because


thin films are used for hard disk media, and other research favors NPs that are
either superparamagnetic – for biomedical applications and ferrofluids, or more
monodisperse – for fundamental studies and self-assembly.
Magnetic nanoparticles can also interact with each other. Quantum mechanical
exchange requires distances of a few Ångstroms. Unless there is an intermediate
magnetic phase to couple the nanoparticles, as in the case of nanocrystalline mate-
rials where an amorphous matrix surrounds Fe-rich nanocrystals [39], exchange
interactions are usually ignored. However, the magnetic moment of a NP generates
a dipolar magnetic field that can be up to tens of kAm−1 nearby. For a grain of
magnetic recording media, this field is small relative to the switching field, but
in composites or assemblies made from NPs with lower anisotropy, magnetostatic
interactions can impact the collective response. There are no simple models of
dipolar interactions in nanoparticle assemblies, since the dipolar field is long-range
and anisotropic. In disordered assemblies or polydisperse samples, this leads to
spin glass-like behavior with a complex energy barrier landscape. Deviations from
noninteracting response are most pronounced in the magnetization dynamics. The
imaginary susceptibility χ  shows differences in ferrofluids with even 1 volume
percent of iron oxide NPs [40]. For the same reason, magnetic hyperthermia
applications where the NPs are exposed to an AC magnetic field will depend on
whether the particles are dilute or densely clustered. In the extreme limit of an
ordered magnetic assembly with strong dipolar interactions, dipolar ferromagnetism
is observed [41, 42]. Here there are multiparticle magnetic domains, and the
magnetization is stable over time. In between the noninteracting and dipolar
ferromagnet cases are superferromagnets, where the moment directions of a cluster
of neighboring particles are correlated, but the direction changes over time [43].
Two opposing effects modify the magnetization curve for interacting NPs [2, 44,
45]. Disorder due to position or size introduces magnetic frustration that leads
to a slower approach to saturation, relative to an assembly of noninteracting
particles. Random crystallographic orientation reduces the effective anisotropy of
the assembly, leading to saturation at reduced fields.
When trying to understand experimental results for magnetic nanoparticles, it is
therefore important to consider how they may differ from the idealized macrospin
model, both in terms of the spin configuration within an individual NP and also in
terms of magnetostatic interactions with other NPs.

Magnetic Nanoparticle Preparation

General Nanoparticle Formation

According to the La Mer model, nanoparticles form when a nucleation threshold


is researched in a supersaturated solution, above which clusters tend to grow rather
than dissolve [46]. This can occur in either the gas or liquid phases. Most magnetic
nanoparticles are synthesized by solution chemistry methods because it is easier to
1018 S. A. Majetich

make the degree of supersaturation spatially homogeneous. Nucleation then occurs


everywhere at the same time, leading to highly monodisperse nanoparticles.
Solid-based routes to magnetic nanoparticles involve either lithography to make
patterned elements, or else ball milling to crush a bulk-like material. Ball milling
introduces structural damage in order to shatter larger particles and can introduce
impurities from the milling material and walls of the container. The particle size
distribution from ball milling is log normal, and this makes it difficult to analyze
the magnetic behavior because a small number of large particles can dominate the
magnetic properties. Nanoparticles made from solids will therefore not be discussed
further in this review.
Gas phase methods have been used but tend to have greater tradeoffs between the
production rate and the particle quality. These particles are as close as possible to
the “bare” NPs of theoretical calculations, but they must be trapped in a passivating
matrix, or they will not stay pristine for long. Cluster beam sources have been used
to measure the evolution of magnetic moment as a function of cluster size [47] and
to prepare NPs of novel materials such as Co2 Si [48] and Fe nitride [49]. While
it is possible to capture the clusters in a matrix of surfactant and then redisperse
them in a carrier solvent, the throughput is low. In contrast, gas phase aerosol
methods can produce kilograms of magnetic NPs. Here the primary particle size is
nanoscale, but the particles are branched or aggregated, and cannot be broken apart.
For a catalyst the large surface area is beneficial, but for ferrofluid, biomedical, or
magnetic storage-related applications, agglomeration is undesirable.

Liquid Phase Syntheses

Liquid phase production of nanoparticles is typically on the milligram to gram scale.


Liquids enable greater control of precursor homogeneity, though high dilution is still
required to prevent coalescence of the growing clusters. Two methods are used to
minimize the likelihood that colliding clusters will stick together: surface charge
for Coulombic repulsion and organic surfactants for steric repulsion. The preferred
synthetic route depends on the end use of the nanoparticles, and details specific
to particular applications will be discussed later. Figure 3 shows the transmission
electron microscopy images of a variety of NPs made by liquid phase synthesis.

Nonaqueous Syntheses
For studies of fundamental physics where monodispersity and high crystallinity are
important, hydrophobic syntheses have been most promising. Here a high boiling
point organic solvent is used to control the thermal decomposition of organometallic
precursors and therefore the nucleation threshold. This is done in a dilute dispersion
in the presence of surfactant molecules to prevent coalescence. The most commonly
used surfactant for magnetic nanoparticles is oleic acid, which has a carboxylic acid
group that binds to magnetic metal atoms at the NP surface. With Au and CdSe
NPs, surfactants with thiol groups are used because they form strong bonds to Au
or Cd. With magnetic nanoparticles there are no surfactants that form such strong
20 Magnetic Nanoparticles 1019

Fig. 3 Transmission electron microscopy images of different types of Fe3 O4 nanoparticles. (a)
Spherical NPs with a size of 13 nm prepared by high-temperature decomposition in nonaqueous
solvent, (b) cube-shaped NPs, (c) NPs made by aqueous coprecipitation, and (d) magnetic beads
containing NPs. (Note: distortion of the beads is due to particle melting of the polystyrene matrix
by the electron beam)

bonds, and there will be an equilibrium between adsorbed and desorbed surfactant
on the NP surface. For this reason care should be taken when diluting a dispersion of
magnetic NPs with a pure solvent. The NP size can be controlled to some extent by
temperature, reaction time, and the precursor to surfactant ratio [50–52]. However,
most of these syntheses have a sweet spot, and it is difficult to find a single synthesis
that can be tuned over a wider size range. The upper limit is often set by the density
of the compound or alloy. Using the same surfactant molecules, it is more difficult to
sterically stabilize dense materials (e.g., Au or FePt) than lighter ones (e.g., Fe3 O4
or γ-Fe2 O3 ). Another problem is the occurrence of antiphase boundaries. Some of
the iron oxide syntheses proceed through a FeO (wustite) NP precursor that is then
transformed into Fe3 O4 . In larger FeO NPs, there can be multiple nucleation sites,
which can lead to antiphase boundaries when the grains meet [25]. The Fe-O-Fe
superexchange across the antiphase boundary is antiferromagnetic and can make
the NPs multi-domain even when they are very small. This can lead to greatly
1020 S. A. Majetich

reduced magnetic moments, and apparently higher effective anisotropy, as seen in


the blocking temperature.

Aqueous Syntheses
Most magnetic NPs for biomedicine are spinel oxide NPs such as Fe3 O4 or
γ-Fe2 O3 made in aqueous solution [53–56]. In this coprecipitation method, metal
salts are dissolved, and a base such as ammonium hydroxide is used to generate
a metal oxyhydroxide phase that loses water to form a spinel ferrite. The reaction
temperature limits the crystallinity of the NPs, though improvements are achieved
by calcining, heating near the boiling point of water, or through hydrothermal
synthesis where the particles are heated under pressure in an autoclave. Once the
NPs are formed, a surfactant is usually added in order to aid in stabilization. Since
iron oxide is amphoteric, collisions between uncoated particles can neutralize the
surface charge. Without surfactant coating a typical charge-stabilized dispersion
will show noticeable settling within a day. Compared with nonaqueous solution
syntheses , the coprecipitation and hydrothermal methods yield particles with
greater polydispersity and lower crystallinity. Depending on the method and particle
size, the specific saturation magnetization of these particles can be significantly
lower than that of the bulk iron oxide.

Nanoparticle Coatings and Extra Requirements of Biomedical


Applications
The NP surface coating prevents agglomeration and helps stabilize the particle
dispersion, since the coating is made of lower density material than the core. To
disperse in nonpolar solvents, the surfactant coating molecules should have nonpolar
tail groups that extend into the solvent. Using dynamic light scattering, it is possible
to measure the hydrodynamic diameter of the NPs and their coating and therefore
to determine whether the particles have single or multiple inorganic cores.
For virtually all biomedical uses, the NPs must form stable aqueous dispersions,
often with a high salt concentration such as that found in blood. There are numerous
ways to modify the as-made NP coating so that they can form aqueous dispersions
[57, 58].
Ex vivo biomedical applications are less restrictive than those where particles
enter the body. NPs must still be chemically inert and biocompatible and form
stable dispersions, but here there is more flexibility to use materials with enhanced
magnetic properties relative to those of iron oxide. A popular method to increase the
magnetic moment while retaining dispersion stability is to encase many magnetic
NPs, usually iron oxide, into a polymer bead that is typically 1–10 microns in
diameter [59]. There are areas for improvement in the magnetic beads. One would
be to improve the magnetic uniformity of the beads, so that they all have the same
net magnetic moment. Ideally this is done with a high packing density, which not
only increases the magnetic moment but also leads to superferromagnetism when the
individual monodomain NPs have strong magnetostatic interactions within the bead.
Compared with an individual superparamagnet, a superferromagnet has a larger
magnetization at lower fields; i.e., it approaches saturation more rapidly. Another
20 Magnetic Nanoparticles 1021

strategy would be to replace magnetite or maghemite with monodomain NPs of a


higher magnetization material such as Fe or FeCo. So far the limitation for this has
been the difficulty of chemical stabilization.
NPs that will be used in vivo have the strictest requirements. They must be
nontoxic, biocompatible, biodegradable, noncarcinogenic, non-mutagenic, and not
trigger an immunoresponse. They cannot sediment or form large agglomerates that
could clog arteries (>20 microns), which could potentially cause a stroke or a heart
attack. For good manufacturing practices, all of the chemicals used in the synthesis
and purification should also be nontoxic. The combination of these needs rules
out anything made by ball milling (some particles will be too big), NPs of cobalt
or nickel (carcinogenic), and NPs made by high-temperature decomposition (due
to residual traces of the high boiling point organic solvents). For these reasons
magnetic NPs intended for in vivo use are typically made in aqueous solution.
Typically they have ∼10 nm particle cores and an organic coating, so they form a
stable aqueous dispersion. In order to retain stability in biological media, which has
a high salt concentration, they are sterically stabilized rather than charge stabilized,
as in an aqueous ferrofluid.
The NP surface coating affects how the body responds to the NP. Protein
molecules quickly adsorb, and depending on the charge, this can signal macrophages
to remove the NPs from circulation. It is important for diagnostic and therapeutic
applications that the NPs are not quickly cleared by the reticuloendothelial system
(RES). Because Fe3 O4 will oxidize to γ-Fe2 O3 over time in aqueous dispersions,
there is often a mixture of phases [60]. It is significant if either Fe2+ from a NP
is solubilized, since this can generate reactive oxygen species (ROS) catalytically.
The onset of numerous neurodegenerative diseases is associated with an upset in
the balance of ROS. The US Food and Drug Administration (FDA) has approved
only a few types of iron oxide (magnetite and maghemite), for in vivo use,
under the trade names Resovist, Feridex, and Endorem. The coatings are typically
polysaccharides. Other magnetic materials may have superior magnetic properties,
but getting approval for in vivo use is a long and expensive process. The European
Commission has been active in sponsoring studies related to health consequences of
nanomaterials [61, 62]. There are no simple answers or sweeping generalizations,
partly because so many factors concerning nanoparticles , including their size,
coating, and degree of dispersion as well, and their chemical composition all
contribute to the biological response.

Applications of Magnetic Nanoparticles

Important Magnetic Characteristics

Most applications of magnetic nanoparticles rely on their small size, which enables
them to be dispersed in liquids and moved with a magnetic field. These applications
depend on several magnetic properties, but the most central is the magnetic moment
| −

μ |= Ms V . The magnitude of μ depends on the material the particle is made
1022 S. A. Majetich

of through Ms and the particle volume V. The maximum V can be limited by


the maximum monodomain size (Eq. 1), or the dispersion stability, which can be
problematic for large sizes or if dense coating materials such as Au are used. If
particle clumping is undesirable, the superparamagnetic limit (Eq. 3) restricts the
maximum size, though the effective moment can be increased by embedding many
magnetic nanoparticles in a polymer to create a magnetic bead, which is typically
microns in diameter.
In magnetic resonance imaging (MRI) contrast agents, the spatially inhomoge-
neous magnetic field generated by μ is given by


→ 2μ cos θ μ cos θ
H = 3
r̂ + θ̂ , (7)
r r3

where θ is the angle between the moment μ and the point of interest. This local field
shifts the resonance frequency of nearby water molecules and enables the removal
of the background signal from water molecules so that the response of protons (1H)
in tissue can be seen more clearly.
In magnetic separation, a spatially varying external magnetic field is used to
apply a force on the magnetic moment:

F = (μ · ∇) H (8)

A strong permanent magnet causes a small deflection of a US one dollar bill,


due to iron oxide NPs embedded in the cellulose fibers in certain regions. Magnetic
paper [63], copying machine toner [64], and hydrogel composites [65, 66] have used
magnetic field gradients to manipulate materials containing magnetic nanoparticles.
Table 1 summarizes the magnetic properties that are most important for different
applications. While there are exceptions, this table is intended as a guide to
understand whether NPs should be superparamagnetic (not stable) and whether
performance could potentially be improved by increasing the magnetic moment (to
generate a larger field and to manipulate in a field or field gradient).

Table 1 Important magnetic characteristics for applications of magnetic nanoparticles


Application Generates field Moves in field Magnetically stable
Magnetic separation, Y (for detection) Y N
manipulation
Magnetic hyperthermia N N N (for DC) but Y (for AC
field used)
Magnetic particle imaging Y (for detection) N N
Magnetic resonance Y N N
imaging
Ferrofluids Y (to form chains) Y N
Magnetic recording media Y (for reading) N Y
20 Magnetic Nanoparticles 1023

Magnetic Separation and Manipulation

Here micron-sized magnetic beads are used instead of individually dispersed


nanoparticles because they provide larger magnetic moment together with a reduced
viscous drag force. There are numerous references that focus on methods to
manipulate NPs [67–69] as well as the effect of NP shape [70].
For biomedical sensing applications, the surfaces of the polymer beads can be
functionalized for selective binding to specific molecules or biomarkers, and a
single bead can have different types of binding sites. In most assays the beads
will have two types of selective binding sites, one to bind to the analyte of interest
and the other to bond to the detection area. The beads are mixed into a complex
dispersion, such as blood, where they diffuse and bind with the analyte. They are
then passed near a functionalized surface where they can bind selectively. This
could be as simple as a glass slide with a permanent magnet on the opposite
surface, so that blood cells infected with malarial parasites could be collected for
optical microscopy analysis [71] or as complex as a magnetoresistive sensor coated
with a thin Au coating that is functionalized to bind selectively with the beads
[72, 73]. In the latter case, beads that are not covalently bound are washed off,
in order to improve the signal-to-noise ratio. The magnetoresistance of the sensor
changes depending on the magnetic field generated by the beads. The interaction
between the magnetic NPs and the free layer of the magnetoresistive sensor is
purely magnetostatic, so the beads must be very close. They must also cover a
large fraction of the sensor area, and they must have high magnetic moment, so
that the field from the particle penetrating into the free layer changes the sensor
resistance.
Other types of magnetic manipulation with NPs are moving toward in vivo
application and focus on therapeutics rather than detection. There have been ex
vivo studies of magnetic vascular repair, in which NPs are forced by a radial
magnetic field gradient to collect on the inner surface of an artery [74, 75]. Here
the magnetic NP delivers a gene for transduction into damaged endothelial cells
that line the arteries. If successful this could lead to a noninvasive method to control
cardiovascular disease.
In magnetic drug delivery, a field gradient would be used to guide the magneti-
cally tagged drug, thereby decreasing the dose required with systemic administra-
tion. At this time, magnetically guided targeting remains a challenge for locations
deep within the body. The surface coating must prevent rapid elimination by the
RES, and while monoclonal antibodies could be attached for selective binding,
targeting through general circulation is very inefficient (<1%), and the antibodies are
expensive. Therefore the magnetic NPs are typically injected into or near a tumor
[76]. However, once in location there is potential for magnetically actuated drug
release when the NPs and the drug are embedded in a sponge-like structure [77].
The dense vasculature of a tumor often helps retain the NPs longer than they
would be in healthy tissue, but the retention rate varies. It has been proposed that
magnetic resonance imaging (MRI) could be used to scan patients and measure
the degree of their NP retention in order to determine the appropriate dosage of
1024 S. A. Majetich

chemotherapy drugs [78]. Cancerous tumors often develop hypoxic regions where
chemotherapy drugs do not penetrate as effectively. A novel suggestion has been
to utilize magnetotactic bacteria for targeting [79]. Normally these bacteria use the
earth’s magnetic field to swim to the aerobic-anaerobic boundary in the sediment of
a pond. When magnetotactic bacteria are bound to drug-containing liposomes, 55%
were able to penetrate the hypoxic regions [79].
Finally, there is a mechanical stimulation-based approach to cancer treatment
involving magnetic nanoparticles [80]. Here micron-sized magnetic disks are
introduced into cultures of tumor cells and then exposed to magnetic fields to apply
an AC magnetic torque . In vitro studies show that this increases cell death. This
work is in the early stages and many questions remain about how to extend it to
in vivo application, but clearly there are multiple ways that magnetic particles can
influence cells.

Magnetic Hyperthermia

Magnetic hyperthermia has become the most popular area for research on biomedi-
cal applications of magnetic NPs. Local magnetic hyperthermia was first used over
60 years ago [81], and in recent years there have been ongoing clinical trials,
but most work is in the preclinical stage. When cancerous tissue is exposed to
temperatures of 42–45 ◦ C, chemotherapy and radiation become more effective.
Temperatures above 50 ◦ C can kill cells directly. The challenge is to kill cancerous
cells but not healthy ones, so localized heating is important. Magnetic NPs bound
to a tumor offer a mechanism to achieve localized heating. The area of a hysteresis
loop corresponds to an energy dissipated as heat. When the loop is cycled by an
AC magnetic field, this leads to power dissipation. The NPs used in hyperthermia
are often referred to a superparamagnetic, but as shown in Eq. 3, this classification
depends on the measurement time. NP sizes that are superparamagnetic at nearly
DC frequencies used for magnetometry (τmeas −1 ∼ 0.01 Hz) may be blocked if
measured at AC frequencies typical of hyperthermia (100 kHz–1 MHz). Viewed
in terms of Eq. 6, there will be a blocking frequency that depends on size and
effective anisotropy. The usable frequency range is determined by the requirement
that H0 f ≤ 4.85 × 108 Am−1 s−1 [82], where H0 is the amplitude of the AC
magnetic field. Above this threshold the AC field causes nerve stimulation. When
NPs are dispersed in a liquid, Brownian rotation, where the whole particle rotates,
can also change the magnetic moment direction, so both Néel and Brownian rotation
must be considered. The rotational diffusion relaxation time τ B is given by.

τB = 3 η dhyd /kB T , (9)

where η is the local viscosity and dhyd is the hydrodynamic diameter.


The most common figure of merit for magnetic hyperthermia is the specific
absorption rate (SAR), defined as the absorbed power per mass of NPs. This is the
20 Magnetic Nanoparticles 1025

clinically meaningful parameter. SAR values of nearly 2.5 kWg−1s at 520 kHz with
a field amplitude of 29 kAm−1 have been reported [83].
The Intrinsic loss power (ILP) defined as SAR/fH0 2 is convenient for comparing
results from measurements with different frequencies and field amplitudes, and it
is meaningful for comparing the intrinsic effects of the NPs. The SAR is measured
by calorimetry, but there are numerous pitfalls that can make it difficult to compare
results from different laboratories [84]. Optical thermometers must be used to avoid
eddy current heating that would arise in a thermocouple sensor. Because the thermal
insulation is imperfect, the heating is measured under nonadiabatic conditions, and
the initial slope of the temperature versus time graph should be used. Data should
be collected as a function of NP concentration. While dense assemblies heat faster,
high concentrations of Fe are toxic [85]. Concentrations of a few mg/mL are most
relevant for clinical application.
There have been numerous simulations of magnetic heating based on combi-
nations of Brownian rotation and Néel-Brown or Stoner-Wohlfarth behavior, all
aiming to maximize the SAR. More sophisticated models have considered the roles
of the size distribution, NP concentration, and magnetic field amplitude [86–88].
Ex vivo experiments have been providing new insights about how magnetic
hyperthermia works. In multicore NPs, the strong coupling between nearby cores
impacts the heating [89, 90]. When NPs are internalized by cells, the SAR is reduced
[91, 92], presumably due to increased viscosity within the cell that would reduce
the Brownian contribution to heating. However, another study reported no change
in the AC susceptibility when NPs were inside cells and after cell lysis [93]. It is
possible that after incorporation into magnetosomes within the cell, the NPs cannot
be redispersed even when the cell walls are ruptured. This would be possible if much
of the NP coating were removed.
According to classical models of heat diffusion, the temperature change should
scale with the size of the heat source, and the temperature increase from a single
nanoparticle should be on the order of 10−8 K [94]. Temperature is an equilibrium
property, but magnetic heating is intrinsically nonequilibrium. A single particle may
be ineffective for hyperthermia, but together a concentration of NPs can raise the
local temperature [95]. One of the most intriguing experimental results has been
the observation of increased cell death after exposure to an AC magnetic field even
when no temperature rise is detected [96]. It is the local power deposition, on the
cellular scale, that leads to cell death. The mechanism by which the local heating
triggers biochemical response that leads to cell death is still unknown, but could
be associated with heat shock proteins. Hyperthermia has been shown to control
ion channel and neurons in microorganisms [97]. The mechanism involved remains
unclear, since ion channels can also be controlled by magnetic forces on attached
magnetic beads, without heating [98].
There have been clinical studies of hyperthermia for 15 years [99], but there are
still many challenges. It is far easier to generate a high amplitude magnetic field in a
laboratory rat than deep inside a person. The wavelength of the AC electromagnetic
field radiation is much larger than the size of a person, so it is difficult to focus
1026 S. A. Majetich

the energy on the tumor target. Most of the initial studies have aimed at tumors of
the breast, prostate, neck, and esophagus, where it is possible with the help of a
probe to achieve therapeutic levels of the AC field. The electrical permittivity and
magnetic permeability of fat and muscle tissue are different [100], and the tissue
structure of the patient must be imaged and modeled [101, 102] to optimize the
hyperthermia dosing treatment. The location of the tumor is also important, because
blood perfusion counteracts the effect of local heating. Magnetic hyperthermia has
the potential to transform certain cancers from life-threatening to chronic diseases
requiring periodic treatment. However, there is still a fair way to go before such
treatment becomes routine.

Magnetic Particle Imaging (MPI)

Magnetic particle imaging can detect moving particles in vivo and so can be used
to monitor biological function in real time. Consider the S-shaped magnetization
curve of a superparamagnet. The magnetization M will oscillate in time in response
to a sinusoidally varying magnetic field. When the field amplitude is high, the
magnetization approaches saturation, so M(t) is not sinusoidal, and a Fourier
transform of the wave would have multiple harmonics. If this process is repeated
with a large DC magnetic field together with the AC field, the AC amplitude can
be much lower and still generate higher harmonics. One of these harmonics is used
to detect the signal from the magnetic particles. Here the particles are micron-sized
polymer beads containing superparamagnetic particles. Frame rates of 50 Hz have
been used to image particles passing through a live mouse heart [103].
While MPI is not a commonly used technique, it has great potential. An
intriguing possibility is to combine it with hyperthermia [104]. Here the MPI field
gradients saturate magnetic NPs everywhere except in a small volume. An AC
magnetic field is then applied for hyperthermia, but only NPs that are not already
saturated can respond , thereby localizing the heating. NPs with an SAR of 150 W/g
were able to selectively heat regions spaced by 3 mm, with a concentration of 5 mg
NP per g of tumor.

Contrast Agents for Magnetic Resonance Imaging (MRI)

The nuclear spins of hydrogen atoms act like paramagnets that have a very high
saturation field. Because of this there is only a slight imbalance between spin-up
and spin-down populations even in large magnetic fields. However magnetization
dynamics are sensitive to the imbalance and can be used to distinguish different
chemical and magnetic environments. Magnetic resonance imaging (MRI) is spatial
mapping of nuclear magnetic resonance (NMR). Ferromagnetic resonance of
electron magnetic moments in a ferromagnet typically occurs at GHz frequencies,
but NMR is commonly in the MHz range, with DC magnetic fields of 1.5–3.0 T.
In micromagnetics the Landau-Lifshitz-Gilbert equations is used to describe
20 Magnetic Nanoparticles 1027

magnetization dynamics, but in the medical field, the Bloch-Bloemergen equations


are used:

∂Mz −
→ − → M s − Mz
= −γ M × B = (10)
∂t z T1

and

∂Mx,y −
→ − → Ms − Mx,y
= −γ M × B = . (11)
∂t x,y T2

Here the static magnetic field B0 is applied in the z-direction, and an rf magnetic
field pulse is applied to rotate the magnetic moments by 90◦ . They precess about the
field as they relax back toward B0 . The rate of relaxation is quantified by the longi-
tudinal relaxation time T1 and the transverse relaxation time T2 . The total magnetic
field B is the sum of B0 and the rf field, which are the same everywhere, plus an
effective magnetic field due to moving electrons, which varies spatially depending
on the local H atom environments. Contrast agents are frequently used to enhance
these differences. T1 contrast agents such as Gd diethylaminetriamine pentaacetate
(Gd DTPA) cause a large decrease in the T1 of protons in water molecules due
to the large Gd3+ magnetic moment. Water is abundant, but the signal from water
molecules provides little information about tissue damage. Magnetic NPs containing
Mn2+ , such as Mn ferrite [105, 106] have also demonstrated potential as T1 contrast
agents, as have Fe3 O4 coated with zwitterion Gd3+ complexes [107], but most NP
contrast agents are iron oxide and have greater effect on T2 of water molecular
protons [106]. T2 is associated with dephasing, whereby different moments precess
at slightly different rates. Here the NP increases dephasing because it generates
an additional magnetic field nearby. The figure of merit for contrast agents is the
relaxivity, which is the change in T1 or T2 per concentration of contrast agent.
Gd DTPA is still the dominant contrast agent, but some people, such as dialysis
patients, cannot tolerate it, and here NP contrast agents are a good alternative. The
NPs used as MRI contrast agents are generally small and not part of magnetic beads
so that they can modify the relaxation of many nearby water molecules. The spatial
resolution of MRI depends on the magnetic field strength and is typically ∼50 μm
for a 3 T field.
The body has natural Fe-containing nanoparticles, mainly in the form of ferritin,
an 8 nm antiferromagnetic iron oxyhydroxide core coated with a protein shell [108].
Uncompensated spins and/or spin canting enables magnetic detection. Because of
this, MRI can be used to diagnose diseases such as thalassemia or iron overload
disease [109], thereby avoiding the need for liver biopsies. The MRI scan is not only
noninvasive but also more accurate because the Fe is not distributed homogeneously
within the liver. Magnetite and maghemite nanoparticles have been found in the
brains of Alzheimer’s patients, but the mechanism of their formation is not yet
understood [110]. Though the abundance is not high enough for MRI detection, the
correlation of magnetic nanoparticle formation and a range of neurodegenerative
1028 S. A. Majetich

diseases such as Huntington’s and Parkinson’s is likely to stimulate future research


in this area.
Due to the limitation of materials suitable for in vivo use, there has been relatively
little research to redesign the magnetic component of the particles. An exception
has been the work of Zabow and coworkers, who have studied 2–10 micron disks
separated by a spacer, in order to obtain contrast at specific resonance frequencies,
depending on the spacing and magnetic layer thickness [111]. This research is in a
preliminary stage and has not yet been studied in vivo.

Ferrofluids

While any liquid dispersion of magnetic nanoparticles is technically a ferrofluid,


this term generally refers to fluids where the particles are not separable by a field
gradient. There are no single-component magnetic liquids under ambient conditions,
but the interactions among nanoparticles in a ferrofluid are so strong that it behaves
like a magnetic liquid. An excellent book by Rosensweig [35] describes the physics
of magnetic fluids. The competition between gravitational and magnetostatic energy
leads to characteristic spikes, or hexagonal arrays of spikes, when a permanent
magnet is brough near a true ferrofluid. Figure 4 shows the images of ferrofluid
spiking. The first ferrofluids were prepared by weeks of mechanical grinding,
followed by dispersion in kerosene in the presence of surfactant [112]. Today liquid
based syntheses are used, and the NPs can be dispersed in either nonpolar or aqueous
solvents, though the nonpolar are more stable.
The distinctive properties of a ferrofluid arise from chain formation due to
magnetostatic interactions. These chains have been observed by cryo-transmission
electron microscopy [113]. They are constantly forming and breaking apart within a

Fig. 4 Ferrofluid spikes (a) top view; (b) side view. Here a commercial ferrofluid of iron oxide
nanoparticles in organic solvent is placed in a glass container that is placed on top of a neodymium
iron boron magnet. The density of spikes and their direction roughly follow the magnetic field lines
from the permanent magnet. (Photos courtesy of A. Abdelgawad)
20 Magnetic Nanoparticles 1029

ferrofluid, which gives rise to the collective response, but also the lack of resistance
to a shear force. The particle diameters range from 5 to 20 nm and are most
commonly iron oxide [114]. The NPs are superparamagnetic, so that the moments
will quickly randomize after the removal of an applied field. Typical ferrofluids
have ∼5 volume percent particles, based on the average core size. With much lower
concentrations, the particles are too far away from each other to induce collective
movement of the fluid droplet. At much higher concentrations, the dispersion
becomes unstable.
The earliest application of ferrofluids in 1938 used the Bitter method [7], but with
colloidal dispersions of Co particles [115] instead of a powder, for the observation
of domain walls by depositing ferrofluid on a ferromagnet surface. Here the fringe
field gradient above the domain walls attracts the particles of the ferrofluid. Today
ferrofluids find frequent use in rotary vacuum seals. An external magnet holds
the ferrofluids in place around a rotating shaft, which can then apply torque
without breaking vacuum. Ferrofluids for this application require low vapor pressure
solvents. Because a ferrofluid has a higher overall density than most liquids, it is a
good heat conductor [116]. At high power there is considerable heat generation
in audio speakers due to resistive heating of the voice coils. When the coils are
immersed in a ferrofluid, which is held in place by permanent magnets, the heat can
be carried away more efficiently than by air cooling alone. The density of the fluid
also helps to damp lower frequencies and unwanted resonances.
A new but promising application of ferrofluids is in magnetic cooling through
the magnetocaloric effect, where temperature changes of up to 28 ◦ C have been
achieved [117]. Here the ferrofluid is made of manganese zinc ferrite NPs that have
a Curie temperature slightly above room temperature. An aqueous dispersion of the
particles flows through a closed coil with a heat sink in one region and a magnetic
field gradient in the other. At 300 K the NPs are ferromagnetic, and due to their
size, superparamagnetic, the ferrofluid is attracted by an external permanent magnet.
When heated above Tc , the particles are only paramagnetic and are able to diffuse
away from the magnet. This leads to self-pumping behavior that reduces the energy
required to cycle the ferrofluid coolant between the heat source and heat exchanger.
Given the environmental concerns about fluorinated and chlorinated hydrocarbon
coolants, magnetic cooling could become the major application of ferrofluids in the
future.

Magnetic Recording Media

Originally, magnetic recording media were made of magnetic NPs. Magnetic


tapes and floppy disks were coated with thin films of needle-shaped magnetic
nanoparticles embedded in a polymer matrix. Monodomain magnets were a natural
choice because the maximum coercivity occurs at the largest monodomain size. A
remanent magnetization is important for memory, and the remanence is correlated
with the coercivity.
1030 S. A. Majetich

Modern hard disk drive media are not made of chemically synthesized nanopar-
ticles, but the concept of a monodomain magnet is still relevant. The media
consist of a monolayer of monodomain magnetic grains embedded in a dielectric
matrix. The media are deposited by co-sputtering onto a seed layer chosen to favor
crystallographic orientation that leads to an easy axis perpendicular to the substrate.
The matrix prevents exchange coupling between grains, and the anisotropy is large
enough that magnetostatic interactions have minimal effect on Hc . A bit contains
multiple grains that are ideally magnetized in the same direction. Variations in grain
size and crystallographic or magnetic orientation lead to variations in the magnetic
fringe field generated by the bit that is detected by the read head. The signal-to-noise
ratio is proportional to the square root of the number of grains per bit. In order to
shrink the bit size without compromising performance, the grain size was reduced.
In the 1990s there was concern about reaching the superparamagnetic limit
with further reduction in grain size, which would be at ∼10 nm for hexagonal Co
[118]. The ability to store information for 10 years is estimated using an energy
barrier model based on monodomain NPs [14]. The thermal stability parameter
 = KV/ kB T is a figure of merit and will ideally be ∼60–80. The problem of the
superparamagnetic limit was overcome by shifting to higher anisotropy alloys such
as CoCrPt, used today with an average grain size of 6.5–8 nm, and L10 FePt, which
is used for heat assisted magnetic recording (HAMR). The high magnetocrystalline
anisotropy of L10 FePt could enable grains as small as 4 nm to be used [119].
Exchange bias has been proposed as a method to enhance stability [120], and
exchange-coupled composite media [121], where the grains are coupled to an
underlayer, has been put into practice. Heat-assisted magnetic recording (HAMR)
media [122] are based on sub-10 nm grains of L10 FePt, where chemical and
structural order are important for uniformity in both the switching field and the
Curie temperature.
The energy barrier model of magnetization reversal assumes an attempt fre-
quency related to the Larmor precession rate. There have been proposals for faster
switching [123], and there has been a demonstration of precessional switching
without thermal activation using ps pulses from a synchrotron, which generate large
magnetic fields ∼20 T [124]. However, extending this to use in hard disk drives is
still a long way off.

Nonbiomedical Topics of Recent Interest

L10 FePt Nanoparticles for Magnetic Recording Media

Self-assembled arrays of L10 FePt nanoparticles were once proposed as magnetic


recording media [50]. Not only was the particle size smaller than the average grain
size in granular recording media, but more importantly, it was a uniform 4 nm as
opposed to ranging from 6.5 to 8 nm, and this regularity was expected to improve
the signal-to-noise ratio in reading bits that contain multiple grains. In the original
20 Magnetic Nanoparticles 1031

paper, the nanoparticles were annealed in order to transform the as-made face-
centered cubic (fcc) or A1 phase into the face-centered tetragonal (fct) or L10 phase,
since the surfactant coating decomposes at temperatures below that of the phase
transformation [125]. A coercivity of 480 kAm−1 was achieved, but at the expense
of sintering the 4 nm particles into an irregular assembly. Many researchers have
since investigated FePt nanoparticles, due to the potential technological significance
for magnetic recording. While this potential has not yet been realized, the results of
these studies provide insight about magnetic alloy nanoparticles.
There were two main synthetic approaches. The first sought to reduce the
temperature required to transform the particles into the L10 phase. The addition
of third elements (Cu, Ag, W) to the FePt during synthesis lowered the minimum
temperature needed to observe the L10 phase. However, the particles were irregular
in shape afterwards. The rate of A1 to L10 transformation was increased because
self-segregation of the third element created vacancies within the FePt. The second
approach was to embed the A1 phase NPs in a matrix that was immiscible to Fe
and Pt, such as MgO [126], NaCl [127], or alumina [128], and then heat to the
temperature needed to obtain the L10 phase. Later work applied this idea to prepare
particles that could be redispersed in liquids after electrocatalysis in acid [129].
However, while the coercivity is large these particles are not monodisperse enough
to self-assemble.
It is more challenging to prepare the phase in nanoparticles than in thin films.
The high coercivity L10 phase has a balance of ferromagnetic and antiferromagnetic
interactions [130], and site occupancy disorder disrupts this balance. The surfaces
of the chemically prepared nanoparticles are often Pt-rich [29, 131]. Recent work
using tilt series images from scanning transmission electron microscopy (STEM)
together with 3D reconstruction was able to locate Fe and Pt atomic positions within
a single 8.4 nm Fe0.28 Pt0.72 nanoparticle [132]. The sample had been annealed under
conditions expected to lead to partial transformation from the A1 phase into either
the L10 phase or the chemical ordered Fe3 Pt or FePt3 L12 phases. The analysis
revealed five distinct L12 grains, three L10 grains, and one Pt-rich A1 grain.
The current outlook for magnetic recording media based on self-assembled
arrays of FePt nanoparticles is not promising. In order to self-assemble, the
magnetic interactions along particles must be weak and isotropic, which would
make it difficult to use self-assembly after phase transformation. The A1 to L10
phase transformation problem is solvable. Crystallographic orientation, which is
critical to uniformity in the magnetic media, will not be easy to achieve starting
from surfactant-coated nanoparticles. That said, FePt continues to be a fascinating
system, and FePt nanoparticles may find applications elsewhere, either in catalysis
or in ex vivo biomedical applications.
One of the most important lessons of the extensive research on FePt NPs is
that even a binary alloy can be hard to synthesize with high homogeneity and
monodispersity. Materials that are nontrivial to prepare in bulk or as thin films,
such as Heusler compounds or Nd2 Fe14 B, are going to be much more challenging
to make in nanoparticle form, if milligram quantities are desired.
1032 S. A. Majetich

Core-Shell NPs

By core-shell NPs we refer to structures where both the core and shell are magnetic
phases. When both the core and shell are magnetic and in direct contact, exchange
interactions dominate, which leads to strong coupling. For a more extended
discussion, readers are encouraged to consult a focused review on this topic [133].
Here we will summarize the main ideas of exchange bias in NPs and highlight some
recent results.
The first core-shell NPs studied had Co cores and CoO shells [134]. Air oxidation
nucleates CoO formation at different locations on the surface, so the shell is
polycrystalline, and there is a range of crystallographic relationships between the
core and shell phases. Since metal atoms migrate to the surface to react with
adsorbed oxygen due to the Kirkendall effect, in extreme cases this can lead to
partly hollow particles [17]. Theoretically a hollow shell could have a vortex rather
than single-domain structure [135], but this has not yet been observed, presumably
because of roughness of structural inhomogeneity in the shell that creates pinning
sites. FePt/MnO dumbbells start out as core-shell NPs, but upon heating phase
segregate into two NPs with a shared interface [136, 137]. Due to exchange bias
effects, they share many of the properties of core-shell NPs.
Most core-shell NPs have a ferro- or ferri-magnetic core and an antiferromag-
netic shell, where the anisotropy of the AF phase is greater than that of the FM
or ferrimagnetic (FiM) phase. In this case an exchange bias effect can be seen in
hysteresis loops, whereby there is a loop shift after cooling in a large magnetic
field, together with a higher coercivity than in the FM core alone. Because of
these features, core-shell NPs have been proposed for use in magnetic recording
media [120]. However, several factors are limiting, including the orientation issue
described for FePt. Specific to exchange-bias systems is the requirement for some
sort of training of the system. This could involve field cooling through the Néel
temperature TN of the AF or exposure of the system to a large magnetic field for
a long time. This is theoretically possible but would require a core-shell NP with a
very high anisotropy AF that has a high TN . However, thermal fluctuations will tend
to relax the ordering faster in a NP than in a thin film. In real particles interfacial
roughness would lead to a broad range of exchange bias fields [138].
Since there are multiple magnetic oxide phases, it is possible to prepare core-
shell NPs with two oxide phases. Of particular interest are Fe3 O4 /γ-Fe2 O3 , which
are both ferrimagnetic, and MnO/Mn3 O4 , where the core is AF below TN = 118 K
and the shell is ferrimagnetic below 43 K [139–141]. In both cases the two phases
have the same structure and differ only in the vacancy concentration. It is therefore
feasible to have a single crystal core-shell particle with epitaxial interfaces. While
small regions with epitaxy have been seen for MnO/Mn3 O4 [141], these do not
extend throughout the NPs. The MnO lattice can be transformed into Mn3 O4 by
cation migration, and this will introduce defects. Low-temperature small angle
neutron scattering (SANS) measurements on ordered assemblies of such NPs do
not show evidence of a uniformly magnetized shell or NP, unlike similar Co/CoO
NPs. It is believed that defects such as antiphase boundaries [25] limit the size and
20 Magnetic Nanoparticles 1033

uniformity of the regions that have exchange bias coupling. With γ-Fe2 O3 , evidence
of gradual transformation of Fe3 O4 to γ-Fe2 O3 is observed by a color change
from black to reddish brown. NPs tailored for biomedical applications, which are
dispersed in water, degrade more rapidly. However, there is no abrupt boundary
between the Fe3 O4 and γ-Fe2 O3 phases, suggesting a gradation of the vacancy
concentration. Most “magnetite-like” NPs have an average vacancy concentration
on the order of 5%. In contrast, there are significant exchange bias effects in NPs
with antiferromagnetic FeO cores and ferromagnetic Fe3 O4 shells [142].
While there have been hundreds of papers about core-shell NPs with exchange
bias between ferro- (or ferri) and antiferromagnetic phases, there are relatively few
on exchange spring NPs with coupling between two ferromagnetic phases with
large and small anisotropy. This is complicated for several reasons. First, many hard
magnetic materials such as Nd2 Fe14 B and Sm2 Co17 have complex crystal structures
that are difficult to achieve by chemical synthesis. From this point of view, L10 FePt
is by far the simplest of the hard magnetic materials, and we have seen that it is
nontrivial to make in NP form. Second, the high oxidation sensitivity of rare earth
materials makes it difficult to find soluble reducing agents. Third, strong exchange
coupling requires lattice matching. Core-shell NPs with Co grown around L10 FePt
have been reported [143]. However, the high coercivity FePt requires coating with
MgO prior to annealing, followed by removal of the MgO, and deposition of the
Co shell. This work builds on earlier research where nanocomposites were prepared
by sintering mixtures of FePt and Fe3 O4 NPs [144]. Until methods are developed
so that large quantities of these NPs could be prepared quickly and economically, it
remains a challenge to apply this approach on a large scale.

Surface Effects and Spin Canting

Even if the NP were made in vacuum and had no chemically distinct coating, the
exchange stiffness and anisotropy would differ from the bulk values because of the
surface. Eq. 1 defines the maximum monodomain size for an ellipsoidal particle
in terms of the saturation magnetization Ms , the effective anisotropy K and the
exchange stiffness A, of the bulk ferromagnetic material. The magnetic moment
per atom has been measured for small Fe, Co, and Ni clusters as a function of size
and shows deviations from the bulk value that oscillate with the cluster size up to
about 500 atoms [47]. These clusters were prepared in a mass-selected cluster beam
and have no surface ligands. However, it is unclear that surface atoms on a larger NP
behave like atoms in a small cluster. Symmetry breaking at the surface contributes
to changes in K in complex ways, depending on whether energy minimization
favors spins perpendicular or parallel to the surface. Since it is impossible for all
spins to be parallel to the surface, the latter condition leads to vortices near the
north and south poles of the NP. In both cases the spin configuration is nonuniform
within the particles unless magnetocrystalline or shape anisotropy dominate surface
contributions. The surface atoms will have fewer magnetic neighbors and therefore
reduced exchange energy per atom leading to lower magnetic exchange stiffness A,
1034 S. A. Majetich

which is proportional to the Curie temperature [8]. The symmetry breaking could
also give rise to Dzhaloshinskii-Moriya contributions to the exchange, so that the
spins need not be parallel. Deviations in Tc have been reported, but interpretation
of experimental results has been tricky. For bulk ferromagnets, Tc is found by
measuring the spontaneous magnetization as a function of temperature in small
magnetic fields and then extrapolating to H = 0, but the spontaneous magnetization
of an ensemble of NPs is zero above the blocking temperature. At TB , spins within
a NP are still coupled, but the moments of different NPs are randomly oriented. At
Tc the spins within a NP are random. Since purely dipolar ferromagnetism has been
observed for NP arrays even close to Tc [42], measurements of the exchange-based
Tc would require measurements on dilute assemblies. This is one possible origin of
the disagreement between reports of increased Tc for assemblies [145, 146] and
decreased Tc for single Ni NPs < 100 nm measured using interference electron
microscopy [147], a technique similar to electron holography. The other possible
artifact arises from deviations in site occupancy. In bulk spinel ferrites, equilibrium
site occupancy is reached after annealing at ∼1000 ◦ C, far higher than any solution
phase synthesis used to prepare NPs.
Most of the nanoparticles discussed here have surface coatings that could
potentially change the magnetic moment of the surface atoms. This effect is
usually small in the case of magnetic oxides such as spinel ferrite (e.g., Fe3 O4 )
NPs coated with organic surfactant molecules such as oleic acid (OA). Here the
carboxylic acid –COO− group of the OA molecule forms two bonds to a surface
Fe2+ or Fe3+ ion, whereas Fe ions in the interior are surrounded by a tetrahedral
or octahedral arrangement of O2− ions. Fe-O-Fe superexchange is responsible
for the ferrimagnetism of spinel ferrites, and there are both ferromagnetic and
antiferromagnetic contributions that depend on the Fe-O-Fe bond angle [148]. Since
OA molecules bind to single surface ions, no additional superexchange pathways are
formed, and the balance is unchanged.
Some of the first evidence for deviations from the macrospin model came
from Mössbauer spectroscopy of spinel ferrite nanoparticles [149, 150]. Mössbauer
spectroscopy is an extremely sensitive probe of the local Fe environment; it can
differentiate the ionic charge, the lattice site, and whether the spin is magnetically
stable (on a 100 ns time scale) or superparamagnetic. In the absence of a magnetic
field, the spectrum for Fe3 O4 nanoparticles can be described by sextet contributions
from Fe3+ ions in tetrahedral (A) sites and Fe2+ and Fe3+ ions in octahedral (B)
sites. For γ-Fe2 O3 nanoparticles, which have no Fe2+ ions, there are two subspectra.
However, in a 5 T field, the same sample has three sextet subspectra, suggesting
that some spins have canted in the high field [150]. The effect has since been
observed in high field Mössbauer spectroscopy for a wide range of spinel ferrite
nanoparticles [151–155]. Surface spins were presumed to be most likely to cant,
but direct evidence came from small-angle neutron scattering, which is sensitive
to the length scale of magnetic order [156]. Using polarization analysis of the
scattered neutrons, three-dimensional magnetic correlations can be differentiated
[157]. Fe3 O4 nanoparticles with a size of 9 nm were found to have uniform
magnetization in zero field, but at moderate fields of 1.4 T, the surface spins formed a
20 Magnetic Nanoparticles 1035

coherent shell that canted coherently. The effect is reversible, and the shell thickness
is temperature-dependent. Subsequent measurements have found minimal distinct
surface canting in CoFe2 O4 NPs that have high K and Tc [158] but strong surface
and core canting in NPs with low K and Tc [159]. A combination of high-resolution
electron energy loss spectrometry and density functional theory calculations has
shown differences in the density of states at the edges of NPs in a high magnetic
field associated with spin canting [160].
While there is ample evidence for spin canting in high fields, there is still debate
about the driving force. Studies of different materials will help to differentiate the
effects of disorder and symmetry breaking. To date spin canting in zero magnetic
field has not been demonstrated, but as characterization techniques improve in
sensitivity and spatial resolution, it may someday be possible to experimentally
verify the predictions of surface anisotropy models [10].

Frontiers and Future Directions

We have surveyed a range of research topics and applications of magnetic nanopar-


ticles. In this final section, we consider state-of-the-art magnetic measurements
and the impact that they may have on future directions of magnetic nanoparticle
research. For magnetism the ultimate limit is the single spin. Single spins have
been measured by magnetic resonance force microscopy [161] and by NV centers in
diamond [162]. Spin-polarized STM has been applied to NP-like magnetic islands,
showing quantum interference effects in Co [163] and antiferromagnetic order
within small fcc Fe islands [164]. These are powerful techniques, but they are not
readily applied to the larger NPs that have been the focus of this review.
The Néel-Brown model has been the basis for understanding magnetization
reversal in ellipsoidal monodomain NPs. It has been verified with measurements
on single 25 nm Co NPs [165, 166], and 20 nm barium ferrite NPs [167] that were
deposited randomly onto a fixed array of microSQUIDs. Scanning SQUIDs have
been used to detect flux from 250 nm patterned Co/Pt multilayers with magnetic
moments of less than 107 μB [168]. Scanning diamond NV centers have been
applied to measurements of bulk YIG [169]. This technique has the sensitivity
needed for single NP measurements. FMR measurements on assemblies of NPs have
very broad resonance peaks [170], primarily due to the different crystallographic
orientations but also from the size and shape. Scanning probe FMR tips have been
demonstrated on YIG thin films [171], and if the sensitivity could be increased,
single NP FMR could be achieved. In a sense, the free layer of a spin torque
oscillator (STO) is a nanoparticle. While this patterned layer is crystallographically
oriented, minute differences in shape have made it difficult to phase lock multiple
STOs.
Magnetoresistance has also been applied to single NP measurements. The first
used a break junctions spin coated with a dilute dispersion of surfactant-coated
Fe3 O4 NPs [32]. As with the microSQUIDs, placement of the NPs relative to the
detector was random, requiring measurement of many devices to identify one with
1036 S. A. Majetich

a NP in the right position. Conductive atomic force microscopy (CAFM) has been
used to measure 18 nm Fe3 O4 NPs deposited on a magnetic thin film that acts as the
reference layer [172]. Here superparamagnetic fluctuations in the magnetoresistance
varied depending on the applied magnetic field. However, there are still no reliable
methods to crystallographically orient the chemically prepared particles, which
would enable single particle measurements of anisotropy. Cube-shaped NPs would
be a natural choice, since they would lie preferentially along a (100) plane.
Magnetic force microscopy (MFM) has been a versatile tool for characterizing
domain structures in thin films, but it relies of the force on the tip magnetic moment
due to the vertical component of the fringe field generated by the sample, and that
has limited its use with NPs. While reduced lift heights can be used, with a spherical
NP, it can be difficult to distinguish topographic and magnetic contrast. To be
detectable by MFM, the NP must be magnetically stable. 715 nm × 148 nm × 50 nm
thick permalloy patterns with magnetic moments ∼10−16 Am2 have been measured
[173], along with 50 nm diameter dots patterned into perpendicular Co/Pd multilay-
ers [174].
Electron microscopy is a common technique for imaging NPs, and magnetic
electron microscopy has been applied to single NPs. Electron holography is sensitive
to the in-plane magnetization and can be used for quantitative measurements, but
it requires a specialized microscope with an electron biprism. Flux from individual
20 nm Co in rings NP has been imaged by electron holography [175], as have chains
of ∼45 nm magnetite NPs that form naturally in magnetotactic bacteria [176]. The
estimated moment of a chain of ∼20 NPs was 7 × 10−16 Am2 , but the features
of individual particles could be seen. Fresnel Lorentz microscopy [41] and electron
holography [42] have been applied to NP arrays and correlated with the conventional
transmission electron microscopy images of the same regions, but single NP features
were not observed. With improvements in instrumentation, together with modeling,
imaging of sub-10 nm NPs is feasible.
In the biomedical area, single magnetic beads have been detected by various
methods. Changes in magnetoresistance due to the attachment of a single magnetic
bead have been demonstrated [177]. While single NPs have been detected by
magnetoresistance [172], this is less likely to translate to biomedical use because
the viscous drag is inversely proportional to the particle diameter. It is easier to
guide a magnetic bead to the attachment site by the sensor. The performance of
magnetoresistive sensing systems is measured in terms of the concentration of
analyte they can detect, rather than the size of the sensor or the number of beads.
Here nonmagnetic factors such as the number of selective binding surface receptors
are likely to improve performance. For MRI contrast from a single bead has been
detected [178]. However, the volume spatial resolution of MRI is still lower than the
volume of a 5 micron bead. MRI can be used to track the location of a bead over
time, but smaller particles would not give better resolution.
In the Néel-Brown model , switching is thermally activated, and the switching
time is proportional to the Larmor precession time. Direct precessional switching
has been shown to be faster [142] but requires a short pulse of a very large magnetic
field. While the initial experiments were done with X-rays at a synchrotron, ultrafast
20 Magnetic Nanoparticles 1037

laser pulses can also lead to rapid switching [179]. Here the magnetization reversal
involves a new mechanism where the size of the magnetic moment changes linearly
over time, rather than remaining constant. Theoretical models predict that short
laser pulses could be used to switch L10 FePt nanoparticles [180], though the linear
reversal mechanism would only apply if there were heated to within a few degrees
of their Curie temperature [181].
In summary, research on magnetic nanoparticles covers many different areas,
both fundamental and applied. Magnetic nanoparticles are an intriguing model
system because they act in many ways like giant spins, though the detailed
spin configurations are more complex. In addition to their relevance to magnetic
recording media, they have found application in medical imaging, diagnostics, and
therapies. Research on magnetic nanoparticles is advancing in different, separate
directions: physics-oriented efforts that focus on the spin configuration, switching,
and magnetization dynamics and interdisciplinary work where the nanoparticles
are a tool that can be manipulated with magnetic fields to understand or control
biological systems.

References
1. Coffey, W.T., Kalmykov, Y.P.: Thermal fluctuations of magnetic nanoparticles: fifty years
after Brown. J. Appl. Phys. 112, 121301 (2012)
2. Majetich, S.A., Sachan, M.: Magnetostatic interactions in magnetic nanoparticle assemblies:
energy, time, and length scales. J. Phys. D. 39, R407–R422 (2006)
3. Pankhurst, Q.A., Connolly, J., Jones, S.K., Dobson, J.: Application of magnetic nanoparticles
in biomedicine. J. Phys. D. 36, R167–R168 (2003)
4. Krishnan, K.M.: Biomedical nanomagnetics: a spin through possibilities in imaging, diagnos-
tics, and therapy. IEEE Trans. Magn. 46, 2523 (2010)
5. Perigo, E.A., Hemery, G., Sandre, O., Ortega, D., Garaio, E., Plazaola, F., Teran, F.J.:
Fundamentals and advances in magnetic hyperthermia. Appl. Phys. Rev. 2, 041302 (2015)
6. Frenkel, J., Dorfman, J.: Spontaneous and induced magnetization in ferromagnetic bodies.
Nature. 126, 274 (1930)
7. Bitter, F.: On inhomogeneities in the magnetization of ferromagnetic materials. Phys. Rev. 38,
1903 (1931)
8. McCurrie, R.A.: Ferromagnetic Materials. Structure and Properties. Academic Press, New
York (1994)
9. Majetich, S.A., Wen, T., Mefford, O.T.: Magnetic nanoparticles. Mater. Res. Soc. Bull. 38,
899 (2013)
10. Berger, L., Labaye, Y., Tamine, M., Coey, J.M.D.: Ferromagnetic nanoparticles with strong
surface anisotropy: spin structures and magnetization processes. Phys. Rev. B. 77, 104431
(2008)
11. Stoner, E.C., Wohlfarth, E.P.: Trans. R. Soc. (London) A. 240, 599 (1948)
12. Brown Jr., W.F.: Thermal fluctuations of a single-domain particle. Phys. Rev. 130, 1677 (1963)
13. Luborsky, F.E.: High coercive materials. J. Appl. Phys. 32, 171S (1961)
14. Sharrock, M.P.: Time-dependent magnetic phenomena and particle-size effects in recording
media. IEEE Trans. Magn. 26, 193 (1990)
15. Néel, L.: Sur les effets des interactions entre les domaines élémentaires ferromagnétiques:
Bascule et reputation. J. Phys. Radium. 20, 215 (1959)
1038 S. A. Majetich

16. Evans, R.F.L., Fan, W.J., Chureemart, P., Ostler, T.A., Ellis, M.O.A., Chantrell, R.W.:
Atomistic spin model simulations of magnetic nanomaterials. J. Phys. Condens. Matter. 26,
103202 (2014)
17. Khurshid, H., Li, W., Phan, M.-H., Mukherjee, P., Hadjipanayis, G.C., Srikanth, H.: Surface
spin disorder and exchange-bias in hollow maghemite nanoparticles. Appl. Phys. Lett. 101,
022403 (2012)
18. Givord, D., Skumryev, V., Nogues, J.: Exchange coupling mechanism for magnetization
reversal and thermal stability of Co nanoparticles embedded in a CoO matrix. J. Magn. Magn.
Mater. 294, 111 (2005)
19. Hoffman, A.: Symmetry driven irreversibilities at ferromagnetic-antiferromagnetic interfaces.
Phys. Rev. Lett. 93, 097203 (2004)
20. O’Grady, K., Fernandez-Outon, L.E., Vallejo-Fernandez, G.: A new paradigm for exchange
bias in polycrystalline thin films. J. Magn. Magn. Mater. 322, 883 (2010)
21. Mazo-Zuluaga, J., Restrepo, J., Munoz, F., Mejia-Lopez, J.: Surface anisotropy, hysterestic,
and magnetic properties of magnetite nanoparticles: a simulation study. J. Appl. Phys. 105,
123907 (2009)
22. De Biasi, E., Ramos, C.A., Zysler, R.D., Romero, H.: Large surface magnetic contribution in
amorphous ferromagnetic nanoparticles. Phys. Rev. B. 65, 144416 (2002)
23. Yanes, R., Chubykalo-Fesenko, O., Kachhachi, H., Garanin, D.A., Evans, R., Chantrell,
R.W.: Effective anisotropies and energy barriers of magnetic nanoparticles with Neel surface
anisotropy. Phys. Rev. B. 76, 064416 (2007)
24. Sun, S., Zeng, H., Robinson, D.B., Raoux, S., Rice, P.M., Wang, S.X., Li, G.: Monodisperse
MFe2 O4 (M = Fe, Co, Mn) nanoparticles. J. Am. Chem. Soc. 126, 273 (2004)
25. Nedelkoski, Z., Kepaptsoglou, D., Lari, L., Wen, T., Booth, R.A., Oberdick, S.D., Gilks, D.,
Ramasse, Q.M., Evans, R.F.L., Majetich, S.A., Lazarov, V.K.: Origin of reduced magnetiza-
tion and domain formation in small magnetite nanoparticles. Sci. Rep. 7, 45997 (2017)
26. Salafranca, J., Gazquez, J., Perez, N., Labarta, A., Pantelides, S.K., Pennycook, S.J., Batlle,
X., Varela, M.: Surfactant organic molecules restore magnetism in metal-oxide nanoparticle
surfaces. Nano Lett. 12, 2499 (2012)
27. Kodama, R.H., Berkowitz, A.E., McNiff Jr., E.J., Foner, S.: Surface spin disorder in NiFe2 O4
nanoparticles. Phys. Rev. Lett. 77, 394 (1996)
28. Kodama, R.H., Berkowitz, A.E.: Atomic-scale modeling of oxide nanoparticles. Phys. Rev.
B. 59, 6321 (1999)
29. Kovacs, A., Sato, K., Lazarov, V.K., Galindo, P.L., Konno, T.J., Hirotsu, Y.: Direct observation
of a surface induced disordering process in magnetic nanoparticles. Phys. Rev. Lett. 103,
115703 (2009)
30. Khurshid, H., Li, W., Chandra, S., Phan, M.-H., Hadjipanayis, G.C., Mukherjee, P., Srikanth,
H.: Mechanism and controlled growth of shape and size variant core/shell FeO/Fe3 O4
nanoparticles. Nanoscale. 5, 7942 (2013)
31. Unni, M., Uhl, A.M., Savliwala, S., Savitzky, B.H., Dhavalikar, R., Garraud, N., Arnold, D.P.,
Kourkoutis, L.F., Andrews, J.S., Rinaldi, C.: Thermal decomposition synthesis of iron oxide
nanoparticles with diminished magnetic dead layer by controlled addition of oxygen. ACS
Nano. 11, 2284 (2017)
32. Lee, S., Fursina, A., Mayo, J.T., Yavuz, C.T., Colvin, V.L., Sofin, R.G.S., Shvets, I.V.,
Natelson, D.: Electrically driven phase transition in magnetite nanostructures. Nat. Mater.
7, 130 (2008)
33. Poddar, P., Fried, T., Markovich, G.: First order metal-insulator transition and spin-polarized
tunneling in Fe3 O4 nanocrystals. Phys. Rev. B. 65, 172405 (2002)
34. Lee, J., Kwon, S.G., Park, J.-G., Hyeon, T.: Size dependence of metal-insulator transition in
stoichiometric Fe3 O4 nanocrystals. Nano Lett. 15, 4337 (2015)
35. Rosensweig, R.E.: Ferrohydrodynamics. Cambridge University Press, Cambridge (1995)
36. Song, Q., Zhang, Z.J.: Shape control and associated magnetic properties of spinel cobalt
ferrite nanocrystals. J. Am. Chem. Soc. 126, 6164 (2004)
37. Ahniyaz, A., Sakamoto, Y., Bergström, L.: Magnetic field-induced assembly of oriented
superlattices from maghemite nanocubes. Proc. Natl. Acad. Sci. 104, 17570 (2007)
20 Magnetic Nanoparticles 1039

38. Singh, G., Chan, H., Baskin, A., Gelman, E., Repnin, N., Král, P., Klajn, R.: Self-assembly of
magnetite nanocubes into helical superstructures. Science. 345, 1149 (2014)
39. Herzer, G.: Grain size dependence of coercivity and permeability in nanocrystalline ferro-
magnets. IEEE Trans. Magn. 25, 3327 (1989)
40. Zhang, J., Boyd, C., Luo, W.: Two mechanisms and a scaling relation for dynamics in
ferrofluids. Phys. Rev. Lett. 77, 390 (1996)
41. Yamamoto, K., Majetich, S.A., McCartney, M.R., Sachan, M., Yamamuro, S., Hirayama,
T.: Direct visualization of dipolar ferromagnetism and domain structures in Co nanoparticle
monolayers. Appl. Phys. Lett. 93, 082502 (2008)
42. Yamamoto, K., Hogg, C.R., Yamamuro, S., Hirayama, T., Majetich, S.A.: Dipolar ferromag-
netic phase transition in Fe3 O4 nanoparticle arrays. Appl. Phys. Lett. 98, 072509 (2011)
43. Petracic, O., Glatz, A., Kleemann, W.: Models for the magnetic ac susceptibility of granular
superferromagnetic CoFe/Al2 O3 . Phys. Rev. B. 70, 214432 (2004)
44. Farrell, D., Cheng, Y., McCallum, R.W., Majetich, S.A.: Magnetic interactions of iron
nanoparticles in arrays and dilute dispersions. J. Phys. Chem. B. 109, 13409–13419
(2005)
45. Petracic, O., Chen, X., Bedanta, S., Kleemann, W., Sahoo, S., Cardoso, S., Freitas, P.P.:
Collective states of interacting ferromagnetic nanoparticles. J. Magn. Magn. Mater. 300, 192
(2006)
46. La Mer, V.K., Dinegar, R.H.: Theory, production, and mechanism of formation of monodisu-
persed hydrosols. J. Am. Chem. Soc. 72, 4847 (1950)
47. Billas, I.M.L., Chatelain, A., de Heer, W.A.: Magnetism from the atom to the bulk in iron,
cobalt, and nickel clusters. Science. 265, 1682 (1994)
48. Balasubramanian, B., Manchanda, P., Skomski, R., Mukherjee, P., Das, B., Geoge, T.A.,
Hadjipanayis, G.C., Sellmyer, D.J.: Unusual spin correlations in a nanomagnet. Appl. Phys.
Lett. 106, 242401 (2015)
49. Xu, Y.H., Hosein, S., Judy, J.H., Wang, J.P.: Iron nitride nanoparticles by naocluster
deposition. J. Appl. Phys. 97, 10F915 (2005)
50. Sun, S., Murray, C.B., Weller, D., Folks, L., Moser, A.: Monodisperse FePt nanoparticles and
ferromagnetic FePt nanocrystal superlattices. Science. 287, 1989 (2000)
51. Yu, W.W., Falkner, J.C., Yavus, C.T., Colvin, V.L.: Synthesis of monodisperse iron oxide
nanocrystals by decomposition of iron carboxylate salts. Chem. Commun. 20, 2306 (2004)
52. Park, J., An, K., Hwang, Y., Park, J.-G., Noh, H.J., Park, J.-H., Hwang, N.M., Hyeon, T.:
Ultra-large-scale syntheses of monodisperse nanocrysals. Nat. Mater. 3, 891 (2004)
53. Massart, R.: Preparation of aqueous magnetic liquids in alkaline and acidic media. IEEE
Trans. Magn. 17, 1247 (1981)
54. Molday, R.S., Mackenzie, D.: Immunospecific ferromagnetic iron-dextran reagents for the
labeling and magnetic separation of cells. J. Immunol. Methods. 52, 353 (1982)
55. Berry, C.C., Curtis, A.S.G.: Functionalisation of magnetic nanoparticles for applications in
biomedicine. J. Phys. D. Appl. Phys. 36, R198–R206 (2003)
56. Torchilin, V.P.: Targeted pharmaceutical nanocarriers for cancer therapy and imaging. Am.
Assoc. Pharm. Sci. J. 9, E128 (2007)
57. Marciello, M., Connord, V., Veintemillas-Verdaguer, S., Verges, M.A., Carrey, J., Respaud,
M., Serna, C.J., Morales, M.P.: Large scale production of biocompatible magnetite nanocrys-
tals with high saturation magnetization values through green aqueous synthesis. J. Mater.
Chem. B. 1, 5995 (2013)
58. Gutiérrez, L., Costo, R., Grüttner, C., Westphal, F., Gehrke, N., Heinke, D., Fornara, A.,
Pankhurst, Q.A., Johansson, C., Veintemillas-Verdaguera, S., Morales, M.P.: Synthesis meth-
ods to prepare single- and multi-core iron oxide nanoparticles for biomedical applications.
Dalton Trans. 44, 2943 (2015)
59. Zhao, H., Saatchi, K., Hafeli, U.O.: Preparation of biodegradable magnetic microspheres with
poly (lactic acid)-coated magnetite. J. Magn. Magn. Mater. 320, 1356 (2009)
60. da Costa, G.M., Blanco-Andujar, C., De Grave, E., Pankhurst, Q.A.: Magnetic nanoparticles
for in vivo use: a critical assessment of their composition. J. Phys. Chem. B. 118, 11738
(2014)
1040 S. A. Majetich

61. Hofmann-Amtenbrink, M., Grainger, D.W., Hofmann, H.: Nanoparticles in medicine: cur-
rent challenges facing inorganic nanoparticle toxicity assessments and standardizations.
Nanomedicine. 11, 1689 (2015)
62. Wells, J., Kazakova, O., Posth, O., Steinhoff, U., Petronis, S., Bogart, L.K., Southern, P.,
Pankhurt, Q., Johansson, C.: Standardisation of magnetic nanoparticles in liquid suspension.
J. Phys. D. 50, 383003 (2017)
63. Olsson, R.T., Azizi Samir, M.A., Salazar-Alvarez, G., Belova, L., Ström, V., Bergland, L.A.,
Ikkala, O., Nogues, J., Gedde, U.W.: Making flexible magnetic aerogels and stiff magnetic
nanopaper using cellulose nanofibrils as templates. Nat. Nanotechnol. 5, 584 (2010)
64. Ziolo, R.F., Giannelis, E.P., Weinstein, B., O’Horo, M.P., Ganguly, B.N., Mehrotra, V.,
Russell, M.W., Huffman, D.R.: Matrix-mediated synthesis of nanocrystalline γ-Fe2 O3 : a new
optically transparent magnetic material. Science. 257, 219 (1992)
65. Xu, X.L., Majetich, S.A., Asher, S.A.: Mesoscopic monodisperse ferromagnetic colloids
enable magnetically controlled photonic crystals. J. Am. Chem. Soc. 124, 13864 (2002)
66. Ramanujan, R.V., Lao, L.L.: The mechanical behavior of smart magnet-hydrogel composites.
Smart Mater. Struct. 15, 952 (2006)
67. Yellen, B.B., Hovorka, O., Friedman, G.: Arranging matter by magnetic nanoparticle
assemblers. Proc. Natl. Acad. Sci. 102, 8860 (2005)
68. Yellen, B.B., Erb, R.M., Son, H.S., Hewlin Jr., R., Shang, H., Lee, G.U.: Traveling wave
magnetophoresis for high resolution chip based separations. Lab Chip. 7, 1681 (2007)
69. Lim, J.K., Lanni, C., Evarts, E., Lanni, F., Tilton, R.D., Majetich, S.A.: Magnetophoresis of
nanoparticles. ACS Nano. 5, 217–226 (2011)
70. Lim, J.K., Tan, D.X., Lanni, F., Tilton, R.D., Majetich, S.A.: Optical imaging and magne-
tophoresis of nanorods. J. Magn. Magn. Mater. 321, 1557–1562 (2009)
71. Zimmerman, P.A., Thomson, J.M., Fujioka, H., Collins, W.E., Zborowski, M.: Diagnosis of
malaria by magnetic deposition microscopy. Am. J. Trop. Med. Hyg. 74, 568 (2006)
72. Osterfeld, S.J., Yu, H., Gaster, R.S., Caramuta, S., Xu, L., Han, S.-J., Hall, D.W., Wilson,
R.J., Sun, S., White, R.L., Davis, R.W., Pourmand, N., Wang, S.X.: Proc. Natl. Acad. Sci.
105, 20637 (2008)
73. Schotter, J., Shoshi, A., Brueckl, H.: Development of a magnetic lab-on-a-chip for point-of-
care sepsis diagnosis. J. Magn. Magn. Mater. 321, 1671–1675 (2009)
74. Vosen, S., Rieck, S., Heidsieck, A., Mykhaylyk, O., Zimmermann, K., Bloch, W., Eberbeck,
D., Plank, C., Gleich, B., Pfeifer, A., Fleischmann, B.K., Wenzel, D.: Vascular repair by
circumferential cell therapy using magnetic nanoparticles and tailored magnets. ACS Nano.
10, 369 (2016)
75. Polyak, B., Medved, M., Lazareva, N., Steele, L., Patel, T., Rai, A., Rotenberg, M.Y., Wasko,
K., Kohut, A.R., Sensenig, R., Friedman, G.: Magnetic nanoparticle-mediated targeting of
cell therapy in stent stenosis in injured arteries. ACS Nano. 10, 9559 (2016)
76. Dutz, S., Kettering, M., Hilger, I., Mueller, R., Zeisberger, M.: Magnetic multicore nanopar-
ticles for hyperthermia – influence of particle immobilization in tumour tissue on magnetic
nanoparticles. Nanotechnology. 22, 265102 (2011)
77. Shademani, A., Zhang, H., Jackson, J.K., Chiao, M.: Active regulation of on-demand drug
delivery by magnetically triggerable microspouters. Adv. Funct. Mater. 27, 1604558 (2017)
78. Miller, M.A., Gadde, S., Pfirschke, C., Engblom, C., Sprachman, M.M., Kohler, R.H., Yang,
K.S., Laughney, A.M., Wojtkiewicz, G., Kamaly, N., Bhonagiri, S., Pittet, M.J., Farokhzad,
O.C., Weissleder, R.: Predicting therapeutic nanomedicine efficacy using a companion
magnetic resonance imaging nanoparticle. Sci. Transl. Med. 7, 314ra183 (2015)
79. Felfoul, O., Mohammadi, M., Taherkhani, S., de Lanauze, D., Xu, Y.Z., Loghin, D., Essa, S.,
Jancik, S., Houle, D., Lafleur, M., Gaboury, L., Tabrizian, M., Kaou, N., Atkin, M., Vuong,
T., Batist, G., Beauchemin, N., Radzioch, D., Martel, S.: Magneto-aerotactic bacteria deliver
drug-containing nanoliposomes to tumour hypoxic regions. Nat. Nanotechno. 11, 941 (2016)
80. Kim, D.H., Rozhkova, E.A., Ulasov, I., Bader, S.D., Rajh, T., Lesniak, M., Novosad, V.:
Biofunctionalized magnetic vortex microdisks for targeted cancer cell destruction. Nat. Mater.
9, 165 (2010)
20 Magnetic Nanoparticles 1041

81. Gilchrist, R.K., Medal, R., Shorey, W.D., Hanselman, R.C., Parrot, J.C., Taylor, C.B.:
Selective inductive heating of lymph nodes. J. Ann. Surg. 146, 596 (1957)
82. Brezovich, I.A.: Low frequency hyperthermia: capacitive and ferromagnetic thermoseed
methods. Med. Phys. Monogr. 16, 82 (1988)
83. Guardia, P., Di Corato, R., Lartigue, L., Wilhelm, C., Espinosa, A., Garcia-Hernandez,
M., Gazeau, F., Manna, L., Pellegrino, T.: Water-soluble iron oxide nanocubes with high
values of specific absorption rate for cancer cell hyperthermia treatment. ACS Nano. 6, 3080
(2012)
84. Wildeboer, R.R., Southern, P., Pankhurst, Q.A.: On the reliable measurement of specific
absorption rates and intrinsic loss parameters in magnetic hyperthermia materials. J. Phys.
D. Appl. Phys. 47, 495003 (2014)
85. Gomez, H.F., McClafferty, H.H., Flow, D., Brent, J., Dart, R.C.: Prevention of gastrointestinal
iron absorption by chelation from an orally administered premixed deferoxamine/charcoal
slurry. Ann. Emerg. Med. 30, 587 (1997)
86. Raikher, Y.L., Stepanov, V.I.: Physical aspects of magnetic hyperthermia: low-frequency ac
field absorption in a magnetic colloid. J. Magn. Magn. Mater. 368, 421 (2014)
87. Serantes, D., Baldomir, D., Martinez-Boubeta, C., Simeonidis, K., Angelakeris, M., Nativi-
dad, E., Castro, M., Mediano, A., Chen, D.-X., Sanchez, A., Balcells, L., Martınez, B.:
Influence of dipolar interactions on hyperthermia properties of ferromagnetic particles. J.
Appl. Phys. 108, 073918 (2010)
88. Ruta, S., Hovorka, O., Chantrell, R.: Unified model of hyperthermia via hysteresis heating in
systems of interacting magnetic nanoparticles. Sci. Rep. 5, 9090 (2015)
89. Cervadoro, A., Giverso, C., Pande, R., Sarangi, S., Preziosi, L., Wosik, J., Brazdeikis, A.,
Decuzzi, P.: Design maps for the hyperthermic treatment of tumors with superparamagnetic
nanoparticles. PLoS One. 8, e57332 (2013)
90. Dennis, C.L., Krycka, K.L., Borchers, J.A., Desautels, R.D., van Lierop, J., Huls, N.F.,
Jackson, A.J., Gruettner, C., Ivkov, R.: Internal magnetic structure of nanoparticles dominates
time-dependent relaxation processes in a magnetic field. Adv. Funct. Mater. 25, 2300 (2015)
91. Di Corato, R., Espinosa, A., Lartigue, L., Tharaud, M., Chat, S., Pellegrino, T., Menager, C.,
Gazeau, F., Wilhelm, C.: Magnetic hyperthermia efficiency in the cellular environment for
different nanoparticle designs. Biomaterials. 35, 6400 (2014)
92. Etheridge, M.L., Hurley, K.R., Zhang, J., Jeon, S., Ring, H.L., Hogan, C., Haynes, C.L.,
Garwood, M., Bischof, J.C.: Accounting for biological aggregation in heating and imaging of
magnetic nanoparticles. Technology. 2, 214 (2014)
93. Soukup, D., Moise, S., Cespedes, E., Dobson, J., Telling, N.D.: In situ measurement of
magnetization relaxation of internalized nanoparticles in live cells. ACS Nano. 9, 231 (2015)
94. Rabin, Y.: Is intracellular hyperthermia superior to extracellular hyperthermia in the thermal
sense? Int. J. Hyperth. 18, 194 (2002)
95. Eggeman, A., Majetich, S.A., Farrell, D.F., Pankhurst, A.Q.: Size and concentration effects
on high frequency hysteresis of iron oxide nanoparticles. IEEE Trans. Magn. 43, 2451–2453
(2007)
96. Riedinger, A., Guardia, P., Curcio, A., Garcia, M., Cingolani, R., Manna, L., Pellegrino, T.:
Subnanometer local temperature probing and remotely controlled drug release based on azo-
functionalized iron oxide nanoparticles. Nano Lett. 13, 2399 (2013)
97. Huang, H., Delikanli, S., Zeng, H., Ferke, D.N., Pralle, A.: Remote control of ion channels
and neurons through magnetic-field heating of nanoparticles. Nat. Nanotechno. 5, 602 (2008)
98. Dobson, J.: Remote control of cellular behaviour with magnetic nanoparticles. Nat. Nanotech-
nol. 3, 139 (2008)
99. Gneveckow, U., Jordan, A., Scholz, R., Bruss, V., Waldofner, N., Ricke, J., Feussner,
A., Hildebrandt, B., Rau, B., Wust, P.: Description and characterization of the novel
hyperthermia-and thermoablation system MFH 300F for clinical magnetic fluid hyperthermia.
Med. Phys. 31, 1444 (2004)
100. Schwan, H.P.: Interaction of microwave and radio frequency radiation with biological
systems. IEEE Trans. Microw. Theory Tech. 19, 146 (1971)
1042 S. A. Majetich

101. Johnson, C.C., Durney, C.H., Massoudi, H.: Long wavelength electromagnetic power absorp-
tion in prolate spheroid models of man and animals. IEEE Trans. Microw. Theory Tech. 23,
739 (1975)
102. Chatterjee, I., Hagemann, M.J., Gandhi, O.P.: Electromagnetic absorption in a multilayered
slab model of tissue under near-field exposure condition. Bioelectromagnetics. 1, 379
(1980)
103. Gleich, B., Weizenecker, J.: Tomographic imaging using the nonlinear response of magnetic
particles. Nature. 435, 1214 (2005)
104. Hensley, D., Tay, Z.W., Dhavalikar, R., Zheng, B., Goodwill, P., Rinaldi, C., Conolly, S.:
Combining magnetic particle imaging and magnetic fluid hyperthermia in a theranostic
platform. Phys. Med. Biol. 62, 3483 (2017)
105. Kim, B.H., Lee, N., Kim, H., An, K., Park, Y.I., Choi, Y., Shin, K., Lee, Y., Kwon, S.G.,
Na, H.B., Park, J.-G., Ahn, T.Y., Kim, Y.W., Moon, W.K., Choi, S.H., Hyeon, T.: Large-scale
synthesis of uniform and extremely small–sized iron oxide nanoparticles for high-resolution
T1 Magnetic resonance imaging contrast agents. J. Am. Chem. Soc. 133, 12624 (2011)
106. Lee, J.-H., Huh, Y.-M., Jun, Y., Seo, W., Jang, J., Song, H.-T., Kim, S., Cho, E.-J., Yoon,
H.-G., Suh, J.-S., Cheon, J.: Artificially engineered magnetic nanoparticles for ultra-sensitive
molecular imaging. Nat. Med. 13, 95 (2007)
107. Zhou, Z., Wang, L., Chi, X., Bao, J., Yang, L., Zhaom, W., Chen, Z., Chen, X., Gao, J.:
Engineered iron-oxide-based nanoparticles as enhanced T1 contrast agents for efficient tumor
imaging. ACS Nano. 7, 3287 (2013)
108. Gider, S., Awschalom, D.D., Douglas, T., Mann, S.: Classical and quantum magnetic
phenomena in natural and artificial ferritin proteins. Science. 268, 5207 (1995)
109. Pierre, T.G.S., Clark, P.R., Chu-anusom, W., Fleming, A.J., Jeffrey, G.P., Olynyk, J.K.,
Pootrakul, P., Robins, E., Lindeman, R.: Noninvasive measurement and imaging of liver iron
concentrations using proton magnetic resonance. Blood. 105, 855–861 (2005)
110. Castellani, R.J., Moreira, P.I., Liu, G., Dobson, J., Perry, G., Smith, M.A., Zhu, X.: Iron:
the redox-active center of oxidative stress in Alzheimer disease. Neurochem. Res. 32, 1640
(2007)
111. Zabow, G., Dodd, S., Moreland, J., Koretsky, A.: Micro-engineered local field control for
high-sensitivity multispectral MRI. Nature. 453, 1058 (2008)
112. Moskowitz, R., Rosensweig, R.E.: Nonmechanical torque-driven flow of a ferromagnetic fluid
by an electromagnetic field. Appl. Phys. Lett. 11, 301 (1967)
113. Butter, K., Bomans, P.H.H., Frederik, P.M., Vroege, G.J., Philipse, A.P.: Direct observation
of dipolar chains in iron ferrofluids by cryogenis electron microscopy. Nat. Mater. 2, 88
(2003)
114. Torres-Diaz, I., Rinaldi, C.: Recent progress in ferrofluids research: novel applications of
magnetically controllable and tunable fluids. Soft Matter. 10, 8584 (2014)
115. Shliomis, M.I., Lyubimova, T.P., Lyubimov, D.V.: Ferrohydrodynamics: an essay on the
progress of ideas. Chem. Eng. Commun. 67, 275 (1988)
116. Odenbach, S.: Ferrofluids and their applications. MRS Bull. 38, 921 (2013)
117. Chaudhary, V., Wang, Z., Ray, A., Sridhar, I., Ramanujan, R.: Self pumping magnetic cooling.
J. Phys. D. 50, 03LT03 (2017)
118. Lu, P., Charap, S.: High density magnetic recording media design and identification:
susceptibility to thermal decay. IEEE Trans. Magn. 31, 2767 (1995)
119. Weller, D., Moser, A., Folks, L., Best, M.E., Lee, W., Toney, M.F., Schwickert, M., Thiele,
J.-U., Doerner, M.F.: High Ku materials approach to 100 Gbits/in2 . IEEE Trans. Magn. 36,
10–15 (2000)
120. Skumryev, V., Stoyanov, S., Zhang, Y., Hadjipanayis, G., Givord, D., Nogues, J.: Beating the
superparamagnetic limit with exchange bias. Nature. 423, 850–853 (2003)
121. Victora, R.H., Shen, X.: Exchange coupled composite media for perpendicular magnetic
recording. IEEE Trans. Magn. 41, 2828 (2005)
122. Meyer, G., Thiele, J.-U.: FePt for HAMR. Phys. Rev. B. 73, 214438 (2006)
123. Bauer, M., Fassbender, J., Hillebrands, B., Stamps, R.L.: Switching behavior of a stoner
particle beyond the relaxation time limit. Phys. Rev. B. 61, 3410 (2000)
20 Magnetic Nanoparticles 1043

124. Back, C.H., Weller, D., Heidmann, J., Mauri, D., Guarisco, D., Garwin, E.L., Siegmann, H.C.:
Magnetization reversal in ultrashort magnetic field pulses. Phys. Rev. Lett. 81, 3251 (1998)
125. Thomson, T., Lee, S.L., Toney, M.F., Dewhurst, C.D., Ogrin, F.Y., Oates, C.J., Sun, S.:
Agglomeration and sintering in annealed FePt nanoparticle assemblies studied by small angle
neutron scattering and x-ray diffraction. Phys. Rev. B. 72, 064441 (2005)
126. Ding, Y., Majetich, S.A.: Size dependence, nucleation, and phase transformation in FePt
nanoparticles. Appl. Phys. Lett. 87, 022508 (2005)
127. Li, D., Poudyal, N., Nandwana, V., Jin, Z., Elkins, K., Liu, J.P.: Hard magnetic FePt
nanoparticles by salt-matrix annealing. J. Appl. Phys. 99, 08E911 (2006)
128. Johnston-Peck, A.C., Tracy, J.B.: Phase transformation of alumina-coated FePt nanoparticles.
J. Appl. Phys. 111, 07B522 (2012)
129. Li, Q., Wu, L., Wu, G., Dong, S., Lv, H., Zhang, S., Zhu, W., Casimir, A., Zhu, H., Mendoza-
Garcia, A., Sun, S.: New approach to fully ordered fct-FePt nanoparticles for much enhanced
electrocatalysis in acid. Nano Lett. 15, 2468 (2015)
130. Brown, G., Kraczek, B., Janotti, A., Schulthess, T.C., Stocks, G.M., Johnson, D.D.: Com-
petition between ferromagnetism and antiferromagnetism in FePt. Phys. Rev. B. 68, 052405
(2003)
131. Antoniak, C., Spasova, M., Trunova, A., Fauth, K., Wilhelm, F., Rogalev, A., Minar, J., Ebert,
H., Farle, M., Wende, H.: Inhomogeneous alloying in FePt nanoparticles as a reason for
reduced magnetic moments. J. Phys. Condens. Matter. 21, 336002 (2009)
132. Yang, Y., Chen, C.-C., Scott, M.C., Ophus, C., Xu, R., Pryor Jr., A., Wu, L., Sun, F., Theis,
W., Zhou, J., Eisenbach, M., Kent, P.R.C., Sabirianov, R.F., Zeng, H., Ercius, P., Miao, J.:
Deciphering chemical order/disorder and material properties at the single-atom level. Nature.
542, 75 (2017)
133. Lopez-Ortega, A., Estrader, M., Salazar-Alvarez, G., Roca, A.G., Nogues, J.: Applications
of exchange coupled bi-magnetic hard/soft and soft/hard magnetic core/shell nanoparticles.
Phys. Rep. 553, 1 (2015)
134. Meiklejohn, W.H., Bean, C.P.: New magnetic anisotropy. Phys. Rev. 105, 904 (1956)
135. Goll, D., Macke, S., Berkowitz, A.E., Bertram, H.N.: Magnetic ground states and the role of
vortices in ferromagnetic hollow spheres. Physica B. 372, 282 (2006)
136. Schladt, T.D., Graf, T., Köhler, O., Bauer, H., Dietzsch, M., Mertins, J., Branscheid, R., Kolb,
U., Tremel, W.: Synthesis and magnetic properties of FePt@MnO nanoheteroparticles. Chem.
Mater. 24, 525 (2012)
137. Sun, X., Klapper, A., Su, Y., Nemkovski, K., Wildes, A., Bauer, H., Köhler, O., Schilmann,
A., Tremel, W., Petracic, O., Brückel, T.: Magnetism of monomer MnO and heterodimer
FePt@MnO nanoparticles. Phys. Rev. B. 95, 134427 (2017)
138. Evans, R.F.L., Bate, D., Chantrell, R.W., Yanes, R., Chubykalo-Fesenko, O.: Influence of
interfacial roughness on exchange bias in core-shell nanoparticles. Phys. Rev. B. 84, 092404
(2011)
139. Salazar-Alvarez, G., Sort, J., Surinäch, S., Baro, M.D., Nogues, J.: Synthesis and size-
dependent exchange bias in inverted core-shell MnO|Mn3 O4 nanoparticles. J. Am. Chem.
Soc. 129, 9102 (2007)
140. Estrader, M., Lopez-Ortega, A., Estrader, S., Golosovsky, I.V., Salazar-Alvarez, G., Vasi-
lakaki, M., Trohidou, K.N., Varela, M., Stanley, D.C., Sinko, M., Pechan, M.J., Keavney,
D.J., Peiro, F., Surinäch, S., Baro, M.D., Nogues, J.: Robust antiferromagnetic coupling in
hard-soft bi-magnetic core/shell nanoparticles. Nat. Commun. 4, 2960 (2013)
141. Berkowitz, A.E., Rodriguez, G.F., Hong, J.I., An, K., Hyeon, T., Agarwal, N., Smith, D.J.,
Fullerton, E.E.: Antiferromagnetic MnO nanoparticles with ferrimagnetic Mn3 O4 shells:
doubly inverted core-shell system. Phys. Rev. B. 77, 024403 (2008)
142. Wetterskog, E., Tai, C.-W., Grins, J., Bergstrom, L., Salazar-Alvarez, G.: Anomalous
magnetic properties of nanoparticles arising from defect structures: topotaxial oxidation of
Fe1_x O|Fe3_δ O4 core|shell nanocubes to single-phase particles. ACS Nano. 7, 7132 (2013)
143. Liu, F., Zhu, J., Yang, W., Dong, Y., Hou, Y., Zhang, C., Yin, H., Sun, S.: Building
nanocomposite magnets by coating a hard magnetic core with a soft magnetic shell. Angew.
Chem. 126, 2208 (2014)
1044 S. A. Majetich

144. Zeng, H., Li, J., Liu, J.P., Wang, Z.L., Sun, S.: Exchange-coupled nanocomposite magnets by
nanoparticle self-assembly. Nature. 420, 395 (2002)
145. Tang, Z.X., Sorensen, C.M., Klabunde, K.J., Hadjipanayis, G.C.: Size-dependent Curie
temperature in nanoscale MnFe2O4 particles. Phys. Rev. Lett. 67, 3602 (1991)
146. Lin, P.-C., Huang, P.-C., Song, K.-J., Lin, M.-T.: Enhanced Curie temperatures in Fe and Co
magnetic nanoparticle assembly on single-crystalline Al2 O3 /NiAl (100) with normal metal
capping layer. Appl. Phys. Lett. 88, 153117 (2006)
147. Nepijko, S.A., Wiesendanger, R.: Size dependence of the curie temperature of separate nickel
particles studied by interference electron microscopy. Europhys. Lett. 31, 567 (1995)
148. Goodenough, J.B., Loeb, A.L.: Theory of ionic ordering, crystal distortion, and magnetic
exchange due to covalent forces in spinels. Phys. Rev. 98, 391 (1955)
149. Sawatzky, G.A., Van Der Woude, F., Morrish, A.H., Mossbauer, A.H.: Study of several
ferrimagnetic spinels. Phys. Rev. 187, 747–757 (1969)
150. Coey, J.M.D.: Noncollinear spin arrangement in ultrafine ferrimagnetic crystallites. Phys.
Rev. Lett. 27, 1140 (1971)
151. Darbandi, M., Stromberg, F., Landers, J., Reckers, N., Sanyal, B., Keune, W., Wende, H.:
Nanoscale size effect on surface spin canting in iron oxide nanoparticles synthesized by the
microemulsion method. J. Phys. Appl. Phys. 45, 195001 (2012)
152. Ngo, A.T., Bonville, P., Pileni, M.P.: Spin canting and size effects in nanoparticles of
nonstoichiometric cobalt ferrite. J. Appl. Phys. 89, 3370–3376 (2001)
153. Zysler, R.D., Fiorani, D., Testa, A.M., Suber, L., Agostinelli, E., Godinho, M.: Size
dependence of the spin-flop transition in hematite nanoparticles. Phys. Rev. B. 68, 212408
(2003)
154. Marx, J., Huang, H., Salih, K.S.M., Thiel, W.R., Schünemann, V.: Spin canting in ferrite
nanoparticles. Hyperfine Interact. 237, 41 (2016)
155. Skoropata, R., Desautels, D., van Lierop, J.: γ-Fe2 O3 nanoparticle intrinsic magnetism
dependence on iron-ion availability during synthesis. J. Appl. Phys. 105, 07B503 (2009)
156. Krycka, K.L., Borchers, J.A., Borchers, J.A., Ijiri, Y., Chen, W.C., Watson, S.M., Laver, M.,
Gentile, T.R., Harris, S., Dedon, L.R., Rhyne, J.J., Majetich, S.A.: Visualizing core-shell
morphology of structurally uniform magnetite nanoparticles. Phys. Rev. Lett. 104, 207203
(2010)
157. Krycka, K., Borchers, J., Ijiri, Y., Booth, R., Majetich, S.: Polarization-analyzed small-angle
neutron scattering. II. Mathematical angular analysis. J. Appl. Crystallogr. 45, 554 (2012)
158. Hasz, K., Ijiri, Y., Krycka, K.L., Borchers, J.A., Booth, R.A., Oberdick, S.D., Majetich, S.A.:
Particle moment canting in CoFe2 O4 nanoparticles. Phys. Rev. B. 90, 180405(R) (2014)
159. Oberdick, S.D., Abdelgawad, A., Moya, C., Mesbahi-Vasey, S., Kepaptsoglou, D., Lazarov,
V.K., Evans, R.F.L., Meilak, D., Skoropata, E., van Lierop, J., Hunt-Isaak, I., Pan, H.,
Ijiri, Y., Krycka, K.L., Borchers, J.A., Majetich, S.A.: Spin canting across core/shell
Fe3 O4 /Mnx Fe3-x O4 nanoparticles. Sci. Rep. 8, 3425 (2018)
160. Negi, D.S., Sharona, H., Bhat, U., Palchoudhury, S., Gupta, A., Datta, R.: Surface spin
canting in Fe3 O4 and CoFe2 O4 nanoparticles probed by high-resolution electron energy loss
spectroscopy. Phys. Rev. B. 95, 174444 (2017)
161. Rugar, D., Budakian, R., Mamin, H.J., Chui, B.W.: Single spin detection by magnetic
resonance force microscopy. Nature. 430, 329 (2004)
162. Toyli, D.M., de las Casas, C.F., Christle, D.J., Dobrovitski, V.V., Awschalom, D.D.: Fluores-
cence thermometry enhanced by the quantum coherence of single spins in diamond. Proc.
Natl. Acad. Sci. 110, 8417 (2013)
163. Oka, H., Ignatiev, P.A., Wedekind, S., Rodary, G., Niebergall, L., Stepanyuk, V.S., Sander, D.,
Kirschner, J.: Spin-dependent quantum interference within a single magnetic nanostructure.
Science. 327, 843 (2010). – 8 nm side triangular Co island structure within due to interference
20 Magnetic Nanoparticles 1045

164. Kubetzka, A., Ferriani, P., Bode, M., Heinze, S., Bihlmayer, G., von Bergmann, K., Pietsch,
O., Blügel, S., Wiesendanger, R.: “Revealing antiferromagnetic order of the Fe monolayer
on W(001)” spin-polarized scanning tunneling microscopy and first-principles calculations.
Phys. Rev. Lett. 94, 087204 (2005)
165. Wernsdorfer, W., Orozco, E.B., Hasselbach, K., Benoit, A., Barbara, B., Demoncy, N.,
Loiseau, A., Pascard, H., Mailly, D.: Experimental evidence of the Néel-Brown model of
magnetization reversal. Phys. Rev. Lett. 78, 1791 (1997)
166. Bonet, E., Wernsdorfer, W., Barbara, B., Benoit, A., Mailly, D.: Three-dimensional magneti-
zation reversal measurements in nanoparticles. Phys. Rev. Lett. 83, 4188 (1999)
167. Wernsdorfer, W., Thirion, C., Demoncy, N., Pascard, H., Mailly, D., Thiaville, A.: Magneti-
sation reversal by uniform rotation (Stoner-Wohlfarth model) in fcc cobalt nanoparticles. J.
Magn. Magn. Mater. 242, 132 (2002)
168. Kirtley, J.R., Paulius, L., Rosenberg, A.J., Palmstron, J.C., Holland, C.M., Spanton, E.M.,
Schiessl, D., Jermain, C.L., Gibbons, J., Fung, Y.-K.-K., Huber, M.E., Ralph, D.C., Ketchen,
M.B., Gibson Jr., G.W., Moler, K.A.: Scanning SQUID susceptometers with sub-micron
spatial resolution. Rev. Sci. Instrum. 87, 093702 (2016)
169. Du, C., van der Sar, T., Zhou, T.X., Upadhyaya, P., Casola, F., Zhang, H., Onbaslo, M.C.,
Ross, C.A., Walsworth, R.L., Tserkovnyak, Y., Yakoby, A.: Control and local measurement
of the spin chemical potential in a magnetic insulator. Science. 357, 195 (2017)
170. Diehl, M.R., Yu, J.-Y., Heath, J.R., Held, G.A., Doyle, H., Sun, S., Murray, C.B.: Crystalline,
shape, and surface anisotropy in two crystal morphologies of superparamagnetic cobalt
nanoparticles by ferromagnetic resonance. J. Phys. Chem. B. 105, 7913 (2001)
171. Lee, I., Obukhov, Y., Xiang, G., Hauser, A., Yang, F., Banerjee, P., Pelekhov, D.V., Hammel,
P.C.: Nanoscale scanning probe ferromagnetic resonance imaging using localized modes.
Nature. 466, 845 (2010)
172. Piotrowski, S.K., Matty, M.F., Majetich, S.A.: Magnetic fluctuations in individual superpara-
magnetic particles. IEEE Trans. Magn. 50, 2303704 (2014)
173. Lederman, M., Gibson, G.A., Schultz, S.: Observation of thermal switching of a single
ferromagnetic particle. J. Appl. Phys. 73, 6961 (1993)
174. Thomson, T., Hu, G., Terris, B.D.: Intrinsic distribution of magnetic anisotropy in thin films
probed by patterned nanostructures. Phys. Rev. Lett. 96, 257204 (2006)
175. Dunin-Borkowski, R.E., Kasama, T., Wei, A., Tripp, S.L., Hytch, M.J., Snoeck, E., Harrison,
R.J., Putnis, A.: Off-axis electron holograph of magnetic chains, rings, and planar arrays of
magnetic nanoparticles. Microsc. Res. Tech. 64, 390 (2004)
176. Dunin-Borkowski, R.E., McCartney, M.R., Frankel, R.B., Bazyinski, D.A., Posfai, M.,
Buseck, P.R.: Magnetic microstructure of magnetotactic bacteria by electron holography.
Science. 282, 1868 (1998)
177. Gaster, R.S., Hall, D.A., Nielsen, C.H., Osterfeld, S.J., Yu, H., Mach, K.E., Wilson, R.J.,
Murmann, B., Liao, J.C., Gambhir, S.S., Wang, S.X.: Matrix-insensitive protein assays push
the limits of biosensors in medicine. Nat. Med. 15, 1327 (2009)
178. Shapiro, E.M., Skrtic, S., Koretsky, A.P.: Sizing it up: cellular MRI using micron-sized iron
oxide particles. Magn. Reson. Med. 53, 329 (2005)
179. Radu, I., Vahaplar, K., Stamm, C., Kachel, T., Pontius, N., Durr, H.A., Ostler, T.A., Barker,
J., Evans, R.F.L., Chantrell, R.W., Tsukamoto, A., Itoh, A., Kirilyuk, A., Rasing, T., Kimel,
A.V.: Transient ferromagnetic-like state mediating ultrafast reversal of antiferromagnetically
coupled spins. Nature. 472, 205 (2011)
180. Lyberatos, A., Weller, D., Parker, G.J.: Switching time in laser pulse heat-assisted magnetic
recording using L10 -FePt nanoparticles. J. Appl. Phys. 117, 133905 (2015)
181. Kazantseva, N., Hinzke, D., Chantrell, R.W., Nowak, U.: Linear and elliptical magnetization
reversal close to the Curie temperature. Europhys. Lett. 86, 27006 (2009)
1046 S. A. Majetich

Sara A. Majetich works on magnetic nanoparticles and spintron-


ics. Her research focuses on methods to characterize and control
magnetism on the nanoscale, including electron holography to
image domains in dipolar ferromagnets, neutron scattering to
demonstrate surface spin canting, and magnetoresistance mea-
surements on single particles. She is a Fellow of the American
Physical Society and the Institute for Electrical and Electronic
Engineers.
Artificially Engineered Magnetic Materials
21
Christopher H. Marrows

Contents
Nanocomposite materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1048
Thin-film structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1049
Bilayers and multilayers with engineered properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1049
2D electron gases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1056
Nanowires . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1058
Template-grown nanowires . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1058
Edge states in 2D topological insulators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1060
Lithographically-patterned 2D ferromagnetic arrays . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1060
Hysteresis of nano/microdisc arrays . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1060
Artificial spin-ice structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1061
Artificial atoms and the Coulomb blockade . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1064
Coulomb blockade . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1064
Single electron spintronics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1065
Single atom manipulation and measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1068
Quantum corrals and wave function imaging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1069
Magnetism at the atomic scale . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1070
Anisotropy of individual atoms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1070
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1071

Abstract

Metamaterials with emergent magnetic properties can be engineered by artifi-


cial means using either “bottom-up” or “top-down” processes. These include
FINEMET-style soft magnets, magnetic multilayers, nanowires, lithographically
patterned arrays, and atomic-scale structures fabricated using scanning probe

C. H. Marrows ()
School of Physics and Astronomy, University of Leeds, Leeds, United Kingdom
e-mail: c.h.marrows@leeds.ac.uk

© Springer Nature Switzerland AG 2021 1047


J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_22
1048 C. H. Marrows

methods. There is a tendency towards control of the material over a larger number
of dimensions as we progress through this list. These different metameterials
also represent increasingly fine levels of control over the atomic-scale structure
that is to be engineered to yield the desired properties. There is also a trend
from bottom-up to top-down assembly methods as we move through this
series of material form factors, indicating the increasing difficulty of imposing
structure on matter. Nevertheless, the emergent properties become more and
more remarkable, and less easy to achieve in other ways, making the mastery
of these difficulties very rewarding, both scientifically and technologically.

Here we discuss artificially engineered magnetic materials. These may be consid-


ered to be ‘metamaterials’, engineered to have functional properties that do not
occur naturally by combining materials, typically at the nanoscale. The techniques
used to engineer the nanostructure range from bulk processes such a thermal
quenches, through thin film deposition and lithography, to sophisticated scanning
probe methods that build the metamaterials on an atom-by-atom basis.

Nanocomposite materials

Composites are systems where distinct materials are intermixed, whilst maintaining
their own separate identities, in order to provide complementary desired properties.
Nanocomposites are those where this mixture of phases occurs at lengthscales that
are substantially below 1 μm. In the field of permanent magnetism, an example
is the exchange spring magnets (dealt with in Chapter PMM §2.6), where high-
moment and high-coercivity materials are mixed at the nanoscale in order to provide
superior permanent magnet performance. Another example from the field of soft
magnetism are the the FINEMET-type materials, which are nanocomposites of
crystallites of α-FeSi in an amorphous matrix [1].
A typical composition for this alloy is Fe73.5 Cu1 Nb3 Si13.5 B9 : the addition of
Cu provides nucleation centres whilst the role of the Nb is to prevent the growth
of large grains of FeSi. A transmission electron micrograph showing the type of
grain structure possessed by this material is shown in Fig. 1. The structure is
nanoscale grains of α-Fe(Si) in a matrix of amorphous residue. The precipitation
of these grains greatly improves the soft magnetic properties with respect to the
precursor. This is explained through the averaging out of any magnetocrystalline
anisotropy over many grains when the grains are smaller than the exchange length
lex = A/μ0 Ms2 [3], as is expected within random anisotropy models [4].
Comparison of mechanically-alloyed material, which lacks any amorphous content,
with melt-spun amorphous ribbon found poorer soft magnetic properties in the
mechanical alloys, underlining the importance the of the amorphous matrix for
exchange coupling the nanoscale grains [2]. FINEMET materials are discussed in
detail in Chapter SMM §6.3.
21 Artificially Engineered Magnetic Materials 1049

Fig. 1 TEM bright field


image and corresponding
selected area diffraction
pattern of melt-spun
FINEMET ribbon heat treated
at 550◦ C for 1 h. The regions
of dark and bright contrast
correspond to crystallites that
make different Bragg angles
with respect to the electron
beam. [2]

Thin-film structures

By sequentially depositing different materials as thin films, metamaterials with


emergent properties may be constructed – although the engineering of the properties
can only be carried out along one dimension, the growth direction of the stack.

Bilayers and multilayers with engineered properties

The development of technologies in the 1980s for the deposition of high-quality


magnetic metal multilayers by molecular beam epitaxy or sputtering led to an
explosion of activity that still continues today. This is because new properties
emerge at the nanoscale, either because the layer thicknesses become comparable
to or less than characteristic lengthscales for the electrons in the metals, or because
properties that arise at interfaces are no longer swamped by those arising from the
bulk of the material. Here, a selection of the most important emergent magnetic
properties are reviewed. In addition, there are many interesting and technologically
important properties related to the transport phenomena in these systems: both
magnetoresistance effects (discussed in Chapter MTR) and current-driven torques
on the magnetisation (treated in Chapter SPC).

Exchange bias
A layer of ferromagnetic material placed in contact with an antiferromagnet will
be exchange coupled to it, the most remarkable manifestation of which is a shift
1050 C. H. Marrows

Fig. 2 Hysteresis loops at 77


K of oxide-coated cobalt
particles. Solid line curve
results from cooling the
material in a 10 kOe field.
The dashed line curve shows
the loop when cooled in zero
field. Reprinted figure with
permission from [W.H.
Meiklejohn, C.P. Bean, Phys.
Rev. 105, 904 (1957).]
Copyright (1957) by the
American Physical
Society. [5].

in the hysteresis loop along the field axis, breaking the usual M(H) = −M(−H)
symmetry. This effect is known as exchange bias, and was first discovered by
Meiklejohn and Bean six decades ago in partially oxidised Co nanoparticles [5].
The shifted hysteresis loop they observed is reproduced in Fig. 2. Notable reviews
of the exchange bias phenomenon over the intervening years include those by
Nogués and Schuller [6] and Stamps [7]. A review of theory was given by Kiwi
[8]. Whilst long considered a curiosity, the ability of exchange bias to provide a
pinned reference layer in a spin-valve structure [9] means that the effect is now of
enormous technological importance in magnetic recording and MRAM applications
(see Chapter MMY).
A notable feature of the data shown in Fig. 2 is that the loop shift is only
observed when the system is field-cooled. When cooled in zero field, the loop
remains symmetric but broadens. In this case, the cooling matters since CoO has a
Néel temperature TN that is just below room temperature, and so is not magnetically
ordered until it is cooled. This gives an important clue as to how exchange bias
comes about. At the heart of the phenomenon is the fact that the spins in the
ferromagnet are exchange-coupled to the spins in the antiferromagnet across the
interface. When a field is applied to the system, only the ferromagnet can respond
21 Artificially Engineered Magnetic Materials 1051

since the antiferromagnet lacks a net moment, but the ferromagnet will exert a
torque on the antiferromagnet. If its spins are held in place, by magnetocrystalline
anisotropy, for instance, then the ferromagnet will not so easily respond to the field.
The requirement for anisotropy in the antiferromagnet was demonstrated using an
artificial exchange bias system [10]. In order to reverse the ferromagnetic spins
whilst leaving those in the antiferromagnet unaffected, it is necessary to break the
exchange bonds at the interface, This energy cost increases the field required to
reverse the ferromagnet magnetisation. When the field is removed, the ferromagnet
returns to its original direction as this exchange energy is recovered. The effect
in low applied fields is unidirectional (as opposed to the more usual uniaxial)
anisotropy, with the ferromagnetic moment ‘pinned’ to point in a particular direction
in low fields. The offset along the field axis is known as theexchange field Hex , given
by the expression

Jex
Hex = , (1)
μ0 MtFM

where tFM is the ferromagnet thickness. This reveals the 1/tFM dependence that is the
signature of an interfacial effect. The field offset is thus a measure of the interfacial
exchange interaction density Jex . Typical values for μ0 Hex are tens of millitesla,
meaning that Jex is orders of magnitude smaller than the usual interatomic exchange
energies. The reason is that it is reduced to an effective value, either by winding an
in-plane domain wall in the antiferromagnet [11], or by averaging over a domain
structure in the antiferromagnet caused by the inevitable atomic scale imperfections
at the interface [12].
Field-cooling is necessary to set the direction in which the moment is pinned. As
the ferromagnet-antiferromagnet system is cooled through the Néel point TN , the
antiferromagnet spin structure orders onto the uniform spin structure of the saturated
ferromagnet, and pins it in place once the field is removed in the cooled state. If the
ferromagnet is not saturated then the local antiferromagnetic order is determined by
the local magnetisation in the domain state of the ferromagnet, leading to a broader,
but unshifted loop. On heating, Hex decreases and vanishes at a temperature that
is referred to as the blocking temperature TB , which is often lower than TN for
the antiferromagnet in question. The width of the hysteresis loop is usually large
at this point [13], and the coercivity remains enhanced up to TN . The blocking
temperature is the point above which the energy barriers in the antiferromagnet
are no longer large enough to prevent the reversal of the entire spin structure in
that layer, following the reversing ferromagnet. Making these irreversible transitions
with the antiferromagnet increases the hysteresis of the entire system.
Perfect antiferromagnetic order is not required for exchange bias to be observed.
Whilst some canonical systems such as CoO, NiO [14], or L10 PtMn [15] show
simple antiferromagnetic spin structures, solid solution Mn alloys such as FeMn
[16] or IrMn [17] are also widely studied, where the spin structure is non-collinear
and lacks spatial periodicity. Exchange bias has even been observed when a
ferromagnetic is coupled to a true spin glass [18].
1052 C. H. Marrows

Interlayer exchange coupling


The study of the coupling of thin magnetic films through metallic spacers has a
long history. In the early 20th century, Maurain observed that when Cu and then
Fe were electroplated onto an Fe wire, the Fe shell became magnetised in the same
direction as the wire if the Cu spacer was sufficiently thin [19], but was unable to
explain the phenomenon. In the 1960s there were similar observations of long-range
ferromagnetic coupling through spacers between five and a few tens of nanometers
thick [20], although in at least some cases this was shown to be due to pinholes [21].
In 1986, Grünberg et al. saw the first hint of something new. They found that
the for the proper thickness of a Cr spacer (0.8 nm), the magnetisations of two Fe
films were coupled antiferromagnetically [23]. It was then found by Parkin et al.
that there is a series of regions of spacer thickness where this antiferomagnetic
coupling appears, not only in Fe/Cr, but also in Co/Cr and Co/Ru [24]. This
oscillatory indirect exchange coupling was subsequently shown to be a quite general
phenomenon for many different spacer materials [22], with the period of the
oscillation usually being ∼ 1 nm (Fig. 3).
An obvious way to try to construct a theory to explain this oscillatory behaviour
is to draw on the Ruderman-Kittel-Kasuya-Yosida (RKKY) interaction, an indirect
exchange interaction between impurity spins in a non-magnetic metal (see Chapter
EXC). The Fermi gas of the host metal naturally tries to screen the charge of
an impurity. However, the fact that the values of wavevector k available to the
conduction electrons are cut off at the Fermi wavevector kF means that screening
at short lengthscales is not very effective, leading to a real-space Friedel oscillation
in electron density around the impurity site. The period of this decaying oscillation
is π /kF . Analagous physics occurs in screening the spin of an impurity [25], giving
an oscillating spin density. The sign of the coupling to a second spin is therefore
dependent on the distance from the first one. It is to be noted that the magnitude
of π /kF ∼ 0.2 nm in most cases, rather shorter than the observed Λ ∼ 1 nm periods.
This is because the fast Friedel oscillation can only be sampled at discrete points
that correspond to the interatomic distance d in the spacer metal. This leads to an
aliasing effect [26], giving an expression

 −1
kF n
= − , (2)
π d

where n is an integer.
These ideas were first used to explain the interactions in rare earth/Y superlattices
by Yafet [27]. In 4f systems the spins are localised on the atomic sites and the
RKKY model is clearly valid. Bruno and Chappert attempted to use the RKKY
model to describe the situation in transition metal multilayers and found that whilst
the periods of the oscillations were correct, the phase and amplitude could not
be properly reproduced [28]. This is because the RKKY model assumes point-
like spins, and does not properly take account of the delocalised magnetism in 3d
materials.
21 Artificially Engineered Magnetic Materials 1053

Fig. 3 Indirect exchange coupling for various transition metal spacers between ferromagnetic Co
layers. The quantities given are the spacer thickness for the first antiferromagnetic coupling peak
A1 , the magnitude of the coupling at this peak J1 , the width of the first coupling peak ΔA1 , and
the period P of the oscillation, as well as the crystal structure and Wigner-Seitz radius rW−S for
each element, in units of Ångströms. An asterisk indicates that only one antiferromagnetically-
coupled spacer-layer thickness region was observed, so it was not possible to directly determine P.
Reprinted figure with permission from [S.S.P. Parkin, Phys. Rev. Lett. 67, 3598 (1991)] Copyright
(1991) by the American Physical Society. [22].

Another physical picture used to describe this phenomenon was developed by


Edwards et al., who treated the coupling using quantum well states in the spacer
by analogy to the de Haas-van Alphen effect [29]. In its simplest form, this model
assumes that the ferromagnetic material is strong (in the sense that the spin-↑ 3d
band is filled, whilst the spin-↓ band has the same form as that of the spacer metal.
This is a fair description of a system like Co/Ru, for instance. This means that for a
parallel alignment of the two ferromagnet moments, the spin-↑ electrons are free to
travel throughout the whole structure, whilst the the spin-↓ electrons are confined to
the spacer layers, with an associated energy cost due to the Heisenberg uncertainty
principle. When the moments are antiparallel, each spin is confined to the half-space
corresponding to the appropriately magnetised ferromagnet. Calculating which of
1
these two states has the lower energy amounts to comparing the parabolic E 2
density of states for a free-electron 1D system to the staircase-like density of states
1054 C. H. Marrows

Fig. 4 Density of states of


confined spin-↓ holes in a
quantum well with

Density of States
spin-dependent reflection
coefficients (solid line), and
of free electrons in the bulk
(broken line).

Fermi
level

Energy

for the confined system, which zigzags above and below it as each quantised energy
level is passed through (see Fig. 4). These quantum well states have been observed
experimentally in inverse photoemission experiments [30].
Both the RKKY and quantum well pictures, as well as other related calculations,
share some common features: an s-d exchange mechanism, polarisation of the
electron gas in the spacer, and periods set by the extremal points of the spacer Fermi
surface. This commonality was made explicit by Bruno in his unified picture [31],
in which he treated the spacer as being analogous to a Fabry-Perot cavity, in which
electron waves reflect back and forth from the interfaces. Since the interfaces are
with ferromagnets, the reflection coefficients are spin-dependent and so the standing
waves set up in the spacer are spin-polarised. Different models then amount to
different ways of calculating these spin-dependent reflection coefficients. This might
be done ab initio, by treating 2D arrays of point spins (the RKKY model), within a
tight-binding model (the quantum well theory), or even just in a simple free electron
picture [32].
It is remarkable that these coupling oscillations are a quantum interference
phenomenon that is robust enough to be observed at room temperature in disordered,
polycrystalline material. Structures exploiting this effect have found applications as
the reference layers in magnetoresistive recording heads [33], and in the operation
of toggle-mode MRAM cells [34]. They also show promise for ultrafast racetrack
memories [35].

Perpendicular magnetic anisotropy


Néel pointed out more than 60 years ago that the symmetry breaking at the surface
or interface of a ferromagnet will give rise to a contribution to the surface energy
that depends on the orientation of the magnetisation [36]. If this anisotropy is of
the right sign, it can lead to an easy axis that points out of the film planet that can
be strong enough to overcome the thin film shape anisotropy that would otherwise
make the film normal a hard magnetic axis. Among the first structures to be built
combining thin layers of different materials in order to engineer magnetic properties
21 Artificially Engineered Magnetic Materials 1055

were multilayers designed to display this perpendicular magnetic anisotropy , such


as Co/Pd [37].
Subsequently systems combining all three 3d ferromagnets, Fe, Co, and Ni,
with a wide variety of different other metals were studied; a good selection of the
experimental results may be found in Ref. [38]. Two points can be made about
the overall trends. The first is that the details of the interface structure (such as
the crystallographic orientation, or the degree of roughness or intermixing) can
have a marked effect on the value of interface anisotropy, with wide variation in
results in a given materials system. The second is that for Co and Fe the interface
anisotropy usually yields a perpendicular easy axis, whilst for Ni the perpendicular
axis is more often magnetically hard. Even for a perpendicular easy axis the critical
thickness above which the demagnetising effects are large enough to cause a spin
reorientation transition back to the film plane is usually only 1-2 nm, often less,
and so ultrathin magnetic films are required to exploit this effect. One way to
(partially) get around this constraint is to make use of a system such as a Co/Ni
multilayer [39], in which a thicker magnetic layer is constructed from many ultrathin
ones, giving a large interface density and hence a sizeable interface anisotropy.
Interest in MgO-based magnetic tunnel junctions led to the development of ultrathin
perpendicularly magnetised Ta/CoFeB/MgO layers for spin transfer torque-written
MRAM applications [40]. The advantage of perpendicular anisotropy in this case is
that the critical energy barrier that must be overcome by the write current pulse no
longer contains the demagnetising energy.

Dzyaloshinskii-Moriya interaction
Usually magnetic order can be described in terms only of the Heisenberg exchange
interaction between neighbouring spins S, which is usually written as − J(S1 ·S2 ).
This scalar product normally leads to collinear states. (The effects of frustration
can prevent this.) However, when structural inversion symmetry is broken, the
Dzyaloshinkii-Moriya interaction [41, 42] (DMI) can also be present, a form of
anisotropic exchange that has a chiral, vector product form: −D · (S1 ×S2 ). The
magnitude of the vector D expresses the strength of the interaction, whilst its
direction is determined by crystal symmetry rules [43].
Inversion symmetry is broken by an interface, and so it is expected to find an
interfacial DMI at the interfaces of a ferromagnetic layer [44, 45]. As shown in
Fig. 5, the DM vector D lies in-plane in this situation. The DMI is a manifestation
of spin-orbit coupling, which is generally weak in 3d ferromagnets, and so the other
metal at the interface needs to be a heavy one from the 5d series for large effects.
Since this is an interfacial effect, very thin films ( 1 nm) are needed in order for it
to play any appreciable role.
Chiral spin states have now been observed in several different systems. A
monolayer of Mn on a W substrate forming a spin spiral was one of the very first
examples [47]. The DMI creates an effective field at a domain wall [48], so that the
domain wall chirality is fixed by the sign of the DMI. Examples systems where this
has been observed include Fe nanowires on a W substrate [49], Co/Ni multilayers
on Pt or Ir substrates [50], and polycrystalline sputtered Pt/Co/AlOx stacks [51].
1056 C. H. Marrows

S2
S1 D12

Large SOC

Fig. 5 Sketch of a DMI at the interface between a ferromagnetic metal (grey) and a metal with a
strong SOC (blue). The DMI vector D related to the triangle composed of two magnetic sites and
an atom with a large SOC is perpendicular to the plane of the triangle. [46].

The DMI also plays a key role in stabilising topologically non-trivial (impossible
to deform smoothly into the ferromagnetic state) vortex-like spin textures with
quasiparticle properties known as skyrmions [52]. This interfacial form of the
DMI is able to stabilise these objects in ultrathin films [53], and the possibility to
construct complex mutilayers means that their properties can be tailored. As well as
examples where truly nanoscale skyrmions have been observed at low temperatures
[54], larger skyrmions [55, 56, 57], and bubble domains with skyrmion topology
[58] have been observed at room temperature. They have been shown to respond to
currents and so can be moved electrically within spintronic devices [59].

2D electron gases

A two-dimensional carrier gas is one where electrons or holes are confined to a


plane such that they are localised on one dimension but behave as (almost) free
carriers in the other two. The most common case is that the carriers are electron-
like and the system is then referred to as a 2-dimensional electron gas (2DEG).
Although the first example of a 2DEG was in a surface state [60], these are
now typically engineered to manifest themselves at at an interface. The classic
examples are in semiconductor heterostructures, such as those in which the quantum
Hall effect (the inversion layer of a silicon metal-oxide-semiconductor field-effect
transistor [61]) and fractional quantum Hall effect (GaAs/AlGaAs interface [62])
were discovered. Two-dimensional materials such as graphene also contain 2DEGs
and can be of such high electronic quality that they display quantised Hall effects at
room temperature [63]. Whilst these Hall effects are driven by magnetic fields, the
materials themselves are not magnetic.
21 Artificially Engineered Magnetic Materials 1057

Interface states in oxide heterostructures


It was discovered in 2004 that the interface between epilayers of the insula-
tors LaAlO3 and SrTiO3 is conducting [64]. Measurements of the Shubnikov-de
Haas oscillations in high mobility layers confirm that this interface state is two-
dimensional and thus a 2DEG is formed [65]. Superconductivity can be observed
at very low temperatures [66], with the superconducting transition temperature of
∼ 200 mK providing a strict upper limit to the thickness of the superconducting layer
of ∼ 10 nm. Electric field effects allow both the normal [67] and superconducting
[68] states to be driven insulating with a gate voltage.
The presence of magnetic order in these systems is less clear-cut. Measurements
of the magnetoresistance of a LaAlO3 /SrTiO3 2DEG revealed hysteresis below
about 300 mK [69], suggesting magnetic domain reorientation. A Hall signal with
all the hallmarks of the anomalous Hall effect was detected at 2 K [70]. On
the other hand, torque magnetometry measurements indicates ferromagnetism that
persists up to room temperature but is lost below 60 K, where a superconductor-like
phase was found [71]. Scanning SQUID microscopy revealed that the magnetism
is inhomogeneous, appearing as submicron patches in a background that has an
inhomogeneous diamagnetic susceptibility [72], implying phase separation into
magnetic and superconducting regions. Signatures of phase coexistence were also
found in the transport properties [73]. Just as for the conductivity, the magnetism is
only present when the LaAlO3 layer is thicker than a few unit cells [74]. The phase
coexistence of ferromagnetism and superconductivity makes this a rather exotic
system that continues to be studied.

Surface states of 3D topological insulators


Topological insulators are materials where the bulk is insulating but the surfaces (in
3 dimensions) or edges (in two dimensions) are conducting [75]. The important
feature is that these surface states are symmetry protected by particle number
conservation and time reversal symmetry. The reason for their existence is that the
topology of the band structure, characterised by a so-called Z2 invariant [76], in the
material is non-trivial, insofar as it takes a different value from that in the vacuum
or a normal band insulator. Whilst both have band gaps at the Fermi level, the gaps
cannot be smoothly connected at the boundary, and so surface states arise.
Whilst not (usually) possessing real space magnetic order, the band-structure in
k-space is of interest for spintronics. The topologically non-trivial band structure
arises due to a strong spin-orbit coupling that inverts bands of different symmetries
about the gap. The surfaces states then take the form of a Dirac cone where the
carrier spin lies in-plane, locked at right angles to the carrier momentum. The first
3-dimensional topological insulator to be predicted was the common thermoelectric
material Bi1−x Sbx [77], which was experimentally demonstrated to possess the spin-
momentum locked Dirac cone surface state using spin-polarised angle-resolved
photoemission spectroscopy [78]. This spin-momentum locking is of interest for
spintronics since it couples the directions of spin-polarisation and current flow
(Fig. 6(b)), as demonstrated by attaching a ferromagnetic contact to the surface of a
1058 C. H. Marrows

a b

2D topological insulator 3D topological insulator

Fig. 6 Spin-momentum locked boundary states in topological insulators. (a) Schematic of the 1D
edge states in a 2D topological insulator. The red and blue curves represent the edge current with
opposite spin character. (b) Schematic of the 2D surface states in a 3D topological insulator. [80].

Bi2 Se3 crystal and measuring the induced voltage for different directions of current
flow in the crystal’s surface [79].
The ordinary Hall effect requires a magnetic field, whilst the anomalous Hall
effect, observed in ferromagnets, does not. Similarly, whilst 2-dimensional systems
can show a quantum Hall effect under field, a quantum anomalous Hall effect is
possible at zero field in the 2-dimensional surface state of a magnetically doped
topological insulator [81], as first observed in a Cr-doped (Bi,Sb)2 Te3 epilayer [82].

Nanowires

Magnetic nanowires are, broadly speaking, magnetic nanostructures that are much
longer in one dimension than the other two. One way to make such a structure
is to pattern it lithographically from a thin film. Nanowires made by this planar
processing approach are common in spintronic applications and can be fashioned
into networks and used to perform logic operations [83] or used as racetrack
memories [84]. These topics are dealt with in more detail in other chapters.

Template-grown nanowires

The most common non-planar method of magnetic nanowire preparation is to


electrochemically deposit material into a narrow nanopore in a template, as reviewed
by Fert and Piraux [85]. Electrodeposition is needed in order to fill a space with such
a huge aspect ratio. Much early work was done using nuclear track-etched polymer
membranes (often polycarbonate) [86], but electrochemically formed arrays of
pores in alumina have also been used [87]. A schematic is shown in Figure 7.
Typical pore diameters are in the range of a few tens to a few hundreds of nm. In
order to form the deposition electrode, a conducting film (typically Au or, as in the
figure, Cu) is coated onto the back of the membrane. Whilst magnetic nanowires of
21 Artificially Engineered Magnetic Materials 1059

Fig. 7 Schematic of an array 40nm


of multilayered nanowires in Polycarbonate
a nanoporous track-etched
polymer membrane. [85]. Cu layer

Co layer

Cu film

a single material are straightforward to grow by this method [88], it is also possible
to form multilayers by either using a pair of baths [89] or by using a single bath and
potentiostatic control toggling between two values of the overpotential [90], which
provides superior results.
The pore densities in anodic aluminium oxide are typically in the range 102 -103
μm−2 , leading to strong dipolar coupling between the nanowires. Arrays of Fe wires
are good examples of strongly coupled systems [91, 92]. Although the easy axis is
usually along the lengths of the wires as expected, the squareness ratio Mr /Ms is
smaller than unity unless the wire diameter is very small, indicating a non-uniform
reversal mode.
On the other hand, nanoporous polycarbonate membranes can have pore densities
as low as 10−2 μm−2 and hence permit the investigation the properties of practically
isolated magnetic nanowires. Enhanced coercivity and squareness can be achieved
in such arrays for suitably small wire diameters [93, 94]. Ni wires made by this
method reveal properties dominated by the shape anistropy of the wire, whilst the
large magnetocrystalline anisotropy of Co modifies their properties due to the local
easy directions that vary from grain to grain [94].
Multilayer nanowires can display a giant magnetoresistance (GMR), measured
in the current-perpendicular-to-plane geometry, with early examples being Cu/Co-
Ni-Cu alloy superlattices [90] and Co/Cu multilayers [95]. The fact that individual
layers can be very thick in structures prepared by this method allowed models of
the GMR in the limit of structures large compared to the spin diffusion length to be
tested [96]. The fact that superconducting contacts are not required due to the high
aspect ratio of the structure means that this is also a method suitable for studying
the temperature dependence of the GMR [97].
In GMR, magnetic configurations affect current flow across interfaces, whilst the
inverse effect where current flow across interfaces affects magnetic configurations
is the domain of spin-transfer torque. It is natural to expect these effects also
to be present in multilayer nanowires. This effect has been used to tailor the
shape of hysteresis loops using current pulses [98] and to effect current-driven
switching [99].
1060 C. H. Marrows

Edge states in 2D topological insulators

Special examples of nanowires are the 1-dimensional channels that form the edge
states in quantum Hall systems. In a conventional (integer) quantum Hall system,
all the carriers travel in the same direction around the edges and the edge states are
said to be chiral.
A two-dimensional topological insulator will have conducting edges that form
helical edge states, as shown in Fig. 6b. The spin-orbit interaction, which can be
thought of as a momentum-dependent magnetic field, spin-momentum locks the
carriers in the same way as in the 3-dimensional systems discussed above. This
means that carriers of opposite spin circulate around the edge states in opposite
directions, leading to the flow of dissipationless pure spin currents in these 1-
dimensional channels.
First introduced theoretically by Kane and Mele [100] and Bernevig and Zhang
[101], to observe the quantum spin Hall effect experimentally requires specially
engineered heterostructures. The first experimental example was a HgTe quantum
well system [102], and more recently InAs/GaSb coupled quantum wells have
emerged as candidate systems [103, 104].

Lithographically-patterned 2D ferromagnetic arrays

The planar processing techniques of microelectronics manufacturing are well-suited


to patterning magnetic thin films into two-dimensional arrays of nanostructures.

Hysteresis of nano/microdisc arrays

The phase diagram of micro- or nanoscale magnetic discs as their thickness and
diameter is varied was determined by Cowburn et al. [105]. This was done by
measuring the shape of the hysteresis loops of an array of Permalloy discs by
the magneto-optical Kerr effect, with the results shown in Fig. 8. When the discs
are very thin and small they are superparamagnetic and show no coercivity. If the
thickness or diameter is increased a little then ferromagnetism is stabilised within a
single domain state, which has leads to a rather square hysteresis loop. Once the disc
is large enough it displays a wasp-waisted loop with a significantly higher saturation
field. This is due to the formation of a vortex state at remanence, which is only
possible when the disc is large enough that savings in magnetostatic energy in this
flux-closed state are sufficient to compensate for the high exchange energy costs
caused by the large magnetisation gradients around the vortex core.
Those results were obtained on discs spaced sufficiently far apart that they are
non-interacting, and hence reproduce the hysteresis properties of an individual disc.
When the spacing is reduced, magnetostatic coupling between the discs comes into
play. A dramatic consequence is that it can stabilise magnetic order in arrays of
21 Artificially Engineered Magnetic Materials 1061

Fig. 8 Hysteresis loops


measured as a function of
diameter d and thickness t
from circular nanomagnets.
For each loop the horizontal
axis is applied field and the
vertical axis is magnetisation.
Reprinted figure with
permission from [R.P.
Cowburn, D.K. Koltsov, A.O.
Adeyeye, M.E. Welland,
D.M. Tricker, Phys. Rev. Lett.
83, 1042 (1999)] Copyright
(1999) by the American
Physical Society.” [105]

discs that are too small to be thermally stable as individuals [106]. The interactions
between single-domain discs in a chain can also be exploited to perform logic
operations magnetically [107].
The flux-closed nature of the vortex state means that it creates very little lateral
stray field, so interdisc coupling is far weaker than in the case of the single domain
state. Nevertheless, just as non-polar molecules can interact via van der Waals
forces, dynamic displacements of the vortex core from the disc centre will produce
a net dipole moment and hence some stray field. This means that whilst the static
states are largely uncoupled, dynamically coupled states can emerge [108]. These
magnonic vortex crystal states have been confirmed experimentally [109].

Artificial spin-ice structures

Spin ices crystals with the pyrochlore structure, where rare earth ions with large
magnetic moments appear at the vertices of the corner-sharing tetrahedra [110].
They are an example of a geometrically frustrated magnetic system: the crystal
geometry means that it is not possible to satisfy all of the interactions in a given
1062 C. H. Marrows

tetrahedron simultaneously. The best that can be done is to find a state where two
of the moments point into the tetrahedron and two point outwards. This is a perfect
analog the proton displacements occuring according to the Bernal-Fowler ‘ice rules’
[111] of the Ice Ih phase of H2 O, leading to exactly the same statistical mechanics
and the prediction of the same residual Pauling entropy [112], hence the name
“spin ice”. Excitations of pyrochlore systems can be described as separated pairs
of magnetic monopole quasiparticles [113].
It is now possible to study geometrical frustration in a in a wide range of artificial
systems that make use of phenomena such as magnetism, superconductivity , and
colloids [114, 115, 116, 117, 118]. Since they are built using nanotechnology,
the structure of such systems, and the frustrated interactions they embody, can be
designed rather than discovered. In magnetic systems, single domain nanoislands
with a well-defined easy axis are used to create Ising-like macrospins that stand in
for the atomic scale magnetic moments in the pyrochlore systems. Such systems are
known as artificial spin ices [119], in which the details of the magnetic configuration
can be inspected macrospin-by-macrospin using advanced magnetic imaging tech-
niques to reveal the exact microstate of the artificial statistical mechanical system.
It is not possible to replicate the tetrahedral symmetry in a 2-dimensional system,
and so a projection into the plane of the substrate of some sort must be made.
Two different ways to do this lead to the square [115, 120, 121, 122] and kagomé
[114, 123, 124, 125] ice arrays that have been widely studied, illustrated in Fig. 9a
and b, respectively. The square ice array has a long-range ordered antiferromagnetic
ground state with the lowest possible twofold degeneracy that is easily accessible
through annealing [126, 127, 128]. By introducing an offset in height between the
two sublattices of the square ice array the equivalence between the interactions
between the four islands that meet at each vertex can be restored, recovering the
extensive degeneracy that is the hallmark of the spin ice state in the pyrochlores
[129]. Meanwhile the kagomé pattern has stronger frustration and a richer phase
diagram [130, 131]; its phase transitions have been probed by low energy muon
spectroscopy [132].
The first examples of these systems were athermal, since the blocking tempera-
tures of the nanoislands were ∼ 105 K, and so the magnetic microstate was entirely
frozen on any realistic laboratory timescale. In order to relax these systems into a
low energy state, various ac demagnetisation protocols were designed [134, 135].
Whilst reducing the system energy, the ground state was never obtained [120, 121].
Although true thermalisation was not possible, an effective thermodynamics was
possible, with analysis of the microstates allowing the effective temperature [136,
137] and entropy [138] to be obtained.
The discovery that it is possible to perform a one-shot anneal into a well-
thermalised state during sample fabrication [126] was quickly followed up by a
demonstration that it is possible to “melt” that macrospin system at temperatures
below the Curie point of the underlying magnetic material in systems where
the blocking temperature is engineered to be experimentally accessible [139].
Annealing systems into their ground state has been made possible through advances
in the materials used to prepare the nanoisland arrays [127, 128, 140, 141], freezing
21 Artificially Engineered Magnetic Materials 1063

Fig. 9 Artificial spin ice lattices. (a) The square (a) and kagome (b) lattices that have been
extensively investigated. The short-island shakti (c) and long-island shakti (d). The darkened
islands in c,d define a plaquette of the lattice, which is the basis for understanding its topologically-
induced emergent frustration. [133].

a thermalised state to become athermal and stable for imaging. Careful control of the
energy barriers and temperature range studied experimentally has allowed the direct
observation of thermal fluctuations and relaxation as artificial spin ices explore their
energy landscape using time-resolved imaging [142, 143, 144, 145].
The ability to design lattices to embody a particular Hamiltonian means that
it is possible to engineer novel emergent phenomena [146], which in many cases
are absent in ‘naturally-occurring’ systems. Examples include the ‘shakti’ lattice
(Fig. 9), which displays topologically induced emergent frustation [133], the
‘tetris’ lattice, which exhibits an emergent reduction in dimensionality [147], spin
fragmentation in the kagome lattice [148], sublattice decomposition in Penrose
patterns [149], and artificial charge ices that display promise as novel data storage
media [150].
1064 C. H. Marrows

Artificial atoms and the Coulomb blockade

The signature feature of the electronic structure of an atom is its discrete energy
spectrum. By a suitable choice of material and size, nanoscale structures can
also be engineered to display a discrete set of energy levels, causing them to be
dubbed ‘artificial atoms’ [151, 152]. The usual way to probe these energy levels
is through electrical measurements, using tunnel contacts so as not to perturb the
electronic structure within the artificial system. This leads to a phenomenon known
as the Coulomb blockade (CB), which is the central feature of single electron
electronics [153]. When combined with magnetic materials, this leads to the field
of nanospintronics [154, 155], in which spin-polarised electrons are manipulated
one-by-one.

Coulomb blockade

Coulomb interaction effects are always present in electrical devices, but it is at


the nanoscale that devices become small enough that the Coulomb effects can
dominate. Nanometre-scale electrically isolated conducting islands give rise to
low capacitances, ∼ 10−18 F, which consequently give large charging energies
EC = e2 /2C, where C is the capacitance of the nanoparticle. One can attach source
and drain tunnel contacts to add and remove electrons from the island. This classical
electrostatic energy arising from the addition of a single electron can be enough to
suppress of transport through the island, which is known as the Coulomb blockade
(CB). The CB can dominate transport if EC  (kB T, |eV |), i.e. for low temperature
T and low bias voltage V . In that case a single tunnel event to the metallic island
can increase the electrostatic energy of the system sufficiently that the bias voltage
is not great enough to allow another electron to enter the dot from the source
until the original one has departed for the drain [156]. These electron tunnelling
events (known as sequential tunnelling) are necessarily be discrete. The Coulomb
blockade occurs when all four energy differences–for tunnelling on and off the dot
from either the source or drain electrode–are positive. Upon increasing the bias
it is possible to inject further electrons each time eV exceeds an integer multiple
of EC , causing steps in the current voltage-curves I(V ) as more electrons can
simultaneously traverse the island due to the increased applied bias, known as the
Coulomb staircase. It is important to note that this discrete set of energy levels arises
through the classical Coulomb interaction in combination with the quantisation of
charge. This is to be distinguished from those in a ‘quantum dot’, where discrete
energy levels arise through quantum confinement [152].
Single electron systems are often discussed using the ‘orthodox theory’, which
is a simplified Fock space treatment where all the excitations are integrated
incoherently. The orthodox theory was initially described in [157, 153], and [158].
The assumptions made are: the electron energy quantisation is ignored, the electron
tunnelling time is assumed to be negligible in comparison with other time scales,
21 Artificially Engineered Magnetic Materials 1065

and cotunnelling (correlated transfer of more than one electron) events are ignored
[153]. This last assumption is a reasonable one if the tunnel contact conductances
are  e2 /h.
When a gate electrode is added to the island to increase the number of terminals
to three, the device is known as a single electron transistor (SET). This can be used
to apply a voltage Vg to the island to shift its energy levels. The conditions for
the four energies exceeding zero now form four slanted lines in a Vg -V diagram,
surrounding a diamond shaped area. A similar set of conditions pertains to each
charge state, leading to a chain of corner-sharing ‘Coulomb diamonds’.

Single electron spintronics

The orthodox theory has been extended to single-electron devices embedded in


magnetic tunnel junctions [159]. One of the most important predictions is that an
oscillatory dependence of tunnel magnetoresistance (TMR) on junction voltage bias
should be observed in the CB regime [160]. The issue of a long spin lifetime, leading
to spin accumulation on the island [161], is important since it can generate new
effects that go beyond the simple Jullière picture (Ref. [162]), including negative
differential resistance, inverse TMR, and the presence of TMR for islands of non-
magnetic material. Despite the approximations made in the orthodox theory, a good
fit to experimental data is often obtained, as shown in [163] (see Fig. 10), [164],
and [165].

Spin accumulation and spin lifetime


The idea of spin accumulation is central to spintronics. It involves the injection
of non-equilibrium spin density into a nonmagnetic material by passing a spin-
polarised current into it from a ferromagnet. The type of structure discussed here
is almost ideal to see the basic mechanism behind the effect. Consider an island

Fig. 10 V dependence of the


TMR through a Co
nanoparticle. The blue, pink,
and orange curves show the
calculated TMR for τ sf = 1
ns, 10 ns and 150 ns,
respectively. The dotted curve
shows the experimentally
observed TMR. The
calculated curve assuming the
spin-relaxation time of 150 ns
well reproduces the
experimental one. [163].
1066 C. H. Marrows

between ferromagnetic electrodes which have antiparallel moments, and let us


suppose that the tunnel rate is spin-dependent and larger for majority spins. This will
mean that the electrons entering the island from the source electrode predominantly
have majority spin, as they may tunnel in more easily. However, it is harder for
them to leave into the drain, whilst minority spins quickly do so. Majority spins
will therefore accumulate in the dot, and equilibrium will only be restored when the
bias voltage is switched off and the current drains away. This idea has been treated
theoretically in a number of different ways [166, 167, 168]. The spin accumulation
can affect the detailed shape of the steps in the staircase I-V curve [169], and does
so in a way that allows the spin polarisation of a ferromagnetic electrode to be
determined. This was implemented experimentally using a STM-based method to
access a single CoFe magnetic nanocluster [170].
The question of the spin lifetime τ sf on the islands in these structures is
an important one, intimately related to spin accumulation, since spins cannot
accumulate if they can relax before leaving the dot. Experiments reported on Fe dots
[171] and on Au ones [172] seem to satisfy this condition but no attempt was made
to quantify the spin lifetime in their analysis. This was done by Wei et al. [173], who
fabricated lateral structures where a single Al nanograin is connected to Permalloy
leads through alumina tunnel barriers. They found that the change in current ΔI
upon switching from a parallel to an antiparallel state hardly changed beyond a
critical bias voltage. This can be used to infer the spin lifetime, which was estimated
to be ∼ 1 μs, and to scale with the electron-phonon relaxation rate according to the
Elliot-Yafet mechanism [174]. This relatively long lifetime is consistent with Al
being a light element where spin-orbit interactions are weak. This can be compared
with the work in Ref. [175], which suggested τ sf ∼ 10 ns on Au dots from the onset
current of TMR. This is nevertheless long compared to the estimate of 150 ps in
the bulk. Spin lifetime enhancements may also be found in ferromagnetic materials
when going from bulk crystals to nanoparticles. The study of planar Co/AlOx /Al
junctions, with Co nanoparticles embedded in the alumina barrier [163] led to a
very clear observation of the oscillating TMR predicted by theory [160], as shown
in figure 10. The fit to these data requires a long spin lifetime, of the order of 150
ns, which is some 104 times longer than is usually found in bulk Co material in the
diffusive regime. A similar extension of τ sf in CoFe was observed using the STM
technique [170].

Cotunnelling effects
As described above, below the Coulomb gap (EC /kB > T and EC /e > V ), sequential
tunnelling processes are blocked by the Coulomb charging effects. Nevertheless, a
small current may still flow, even in the ideal case of no leakages, by considering
higher order processes [176], the most commonly discussed of which is known as
cotunnelling. It was predicted by Takahashi and Maekawa that in this cotunnelling
regime the TMR can be substantially enhanced when the magnetisation of the
central island is switched against the outer electrodes. This is because this coherent
process leads to the overall resistance of the double arising from the product of
the individual junction resistances, whereas it follows the usual series resistor sum
21 Artificially Engineered Magnetic Materials 1067

rule in the sequential regime. It can be shown within the Jullière model that the
TMR in the cotunnelling regime should be enhanced by a factor of 2/(1 − P2 ),
where P is the electrode spin polarisation. These effects were explored in some
detail using a conventional double junction stack with granular CoFe deposited
between two outer pinned CoFe electrodes [177]. At 7 K the TMR ratio was 24 %
at zero bias, roughly double the value measured beyond the gap at V = 100 mV. The
effective spin polarisation P was estimated to be 32 % from the non-enhanced TMR,
leading to a prediction from the 2/(1 − P2 ) expression of an enhanced TMR of 23 %,
in agreement with the observed value. Another magnetic configuration is double
magnetic tunnel junctions with conventional free and pinned outer electrodes, in
which superparamagnetic NiFe nanoparticles form the central electrode sandwiched
between alumina barriers [178]. Whilst the TMR in the CB regime is a little
lower than that for a control sample lacking the NiFe islands, the cotunnelling
enhancement in the TMR was clearly observed at biases small enough to lie
within the Coulomb gap for the double MTJ. This shows that spin information
can be propagated through the fluctuating NiFe island moments. This is reasonable
since typical tunnelling times are of the order of femtoseconds, much shorter than
typically superparamagnetic fluctuation lifetimes, which are ∼ 0.1-1 ns.

Kondo effect
The Kondo effect arises when the spin of a magnetic impurity interacts with
the surrounding free electrons to form a singlet state. The earliest studies of the
Kondo effect considered the situation of doping a conductor, such as copper, with
a magnetic impurity [179]. As a result, the conductivity saturates when reducing
the temperature. The maximum value of the conductivity is achieved for the so
called Kondo temperature (TK ) [180]. This value directly depends on the number
of defects introduced in the conductor. In single electtron devices, the a net spin
in the artificial atom can interact with the electrons in the leads to form the Kondo
singlet [181]. In tunnel junctions with magnetic nanoclusters in the insulating layer,
the Kondo effect can produce an enhancement of the resistance or its reduction
depending on the exact location of the clusters within the layer. For instance,
Kondo resonances were suggested to be be the cause of a rise in the resistance
at low temperature and low bias when introducing thin Cr(< 0.4 nm)/Co(< 0.6 nm)
impurities in one of the electrode/barrier interfaces of a magnetic tunnel junctions
[182]. On the other hand, the Kondo effect produces an increase of the conductance
if the insulating tunnel barrier is doped with a magnetic material [183] or some
magnetic nanoclusters are placed within this layer [184]. In the CB regime, where
nanoclusters are intentionally placed within the insulator, some theoretical work
[185, 186], and recent experimental results [187], show that the Kondo effect
in TMR can be important even if the magnetic moments have large magnetic
anisotropy [188]. A theoretical discussion of the Kondo effect in single-electron
devices can be found in Ref. [189]. The crossover between Kondo TMR suppression
and co-tunneling enhancement of TMR in double magnetic tunnel junctions with
CoFe nanoclusters within the MgO barrier is discussed in [187]. For all CoFe
thicknesses the tunnelling is dominated by sequential tunnelling at high temperature
1068 C. H. Marrows

and bias voltage. However, when the temperature is reduced, some typical Kondo
signatures can be observed: a zero bias anomaly (ZBA) in the conductance versus
bias voltage curve; the temperature dependence of the ZBA; and suppression of the
TMR at low bias below a particular temperature. From each of these experimental
features, the Kondo temperature can be determined. The crossover from the two
effects, Kondo and CB, was found to be correlated with the fluctuations of the
magnetic moments of the nanoclusters. A related experiment using high-anisotropy
CoPt nanoclusters found a gradual competition between cotunneling enhancement
of the TMR and the TMR suppression due to the Kondo effect, with both effects
having been found to coexist even in the same sample. It is possible to tune between
these two states with temperature [190].

Chemical potential effects


Modifications of the chemical potential μ in nanodots can have a significant effect
on transport through them. This is, after all, the basis of the operation of a gate
electrode in a SET. The magneto-Coulomb effect, discovered by Ono et al. [191,
192] was studied by theoretically [193]. The central issue is that the application
of a magnetic field H can modulate the μ in a ferromagnet through the flux
density B it gives rise to. CB anisotropic magnetoresistance is another effect
related to manipulation of μ in a ferromagnetic dot [194], in this case one with
a large spin-orbit coupling. There, the chemical potential shifts are related to the
uniaxial magnetocrystalline anisotropy, which arises from spin-orbit coupling, and
the magnetisation angle plays the role of a gate voltage. Another related experiment
studied transport through a single Au nanoparticle connected to Co leads [165].
The I(V ) characteristics of this device could be fitted well using the orthodox
theory of CB, with the commonplace requirement of a local charge offset Q0
reflecting the local electrostatic environment of the nanoparticle. Again, this can
be seen as a chemical potential shift μ = Q0 e/C, with C the capacitance of the
dot and e the electronic charge. Remarkably, this charge was found to vary simply
with the direction of magnetisation of the two Co electrodes in a saturating field:
the moment direction is coupled through spin-orbit interactions to the chemical
potential of the dot and hence charge transfer from the leads. A 90◦ rotation
corresponded to a change in background charge ΔQ0 = 0.033e, giving rise to large
conductance changes near the Coulomb steps in the I(V ) curve. As with the other
magneto-Coulomb effects described above, this essentially allow spintronic SETs
to be constructed where the transistor action is gated by a magnetic, rather than an
electric, field.

Single atom manipulation and measurements

Whilst it is a commonplace that a scanning tunnelling microscope (STM) can image


surfaces with atomic resolution, the ability to transfer an single atom from the tip
to a surface [195] and to position individual atoms on a surface [196] mean that it
is a tool that can also be used to manipulate matter with atomic resolution, albeit
21 Artificially Engineered Magnetic Materials 1069

painstakingly. This means that materials can be engineered at the fundamental limit
imposed by the discreteness of matter.

Quantum corrals and wave function imaging

In one of the iconic experiments of nanotechnology, Crommie et al. used these


capabilities of the STM to build ‘quantum corrals’ [197]. In this first experiment,
electrons in the Shockley surface state of (111) copper were confined within a
circular barrier constructed from 48 iron adatoms. These formed a circular corral
of radius 7.13 nm, within which scanning tunnelling spectroscopy revealed that the
two-dimensional wavefunction takes the form expected for a particle in a circular
box, illustrated in Fig. 11.
In the original experiments the magnetic nature of the adatoms was not relevant,
but the fact that a Co adatom possesses a magnetic moment was of critical
importance to another classic experiment in this field: the observation of a ‘quantum
mirage’ by Manoharan et al. [199]. Here the physics of quantum corrals is combined
with the Kondo effect, which describes the many-body response of an electron gas
to a magnetic impurity. In this experiment, a ellipse-shaped corral was formed, with
a single Co atom placed at one of the two foci. The spatially localized spectroscopic
response of the Kondo impurity was found at the other focus, where there was in
fact no Kondo impurity. The mirage experiment achieved this by taking advantage of
both the locally modified electron density in the corral and the scattering properties
of a Kondo impurity. A scattering theory has been developed that is remarkably
successful in reproducing every detail of the experiments including the electron
standing-wave patterns, the energies and widths of corral states, and all features
of the quantum mirage [200].

Fig. 11 An STM image of a


circular quantum corral
constructed from 48 Fe
atoms. The diameter of the
corral is 143 Å. The circular
pattern in the center of the
corral is the density
distribution due to 3 nearly
degenerate quantum states of
the corral. [198]
1070 C. H. Marrows

Magnetism at the atomic scale

The STM can also be used to investigate magnetism at the atomic scale, for
instance by measuring the magnetisation curve of a single atom [201]. Doing so
for individual Fe atoms on a Cu (111) surface, combined with inelastic scanning
tunneling spectroscopy revealed that the typical moment of such an atom is 3.5 Bohr
magnetons with a magnetic excitation lifetime of 200 fs [202]. The findings were
quantitatively explained by the decay of the magnetization excitation into Stoner
modes of the itinerant electron system. Longer timescales of between 50 and 250 ns
were found using STM-based pump-probe methods [203].
Combining these measurement methods with the ability to engineer nanos-
tructures specified atom-by-atom, including the exchange coupling between the
moments on neighbouring atoms (by controlling the interatom distance to yield
the selected value of the oscillatory indirect exchange) between them [201], has
led to the ability to make all-spin logic gates from atomic chains [204], control the
ground states of chains (depending on even or odd numbers of constituent atoms),
engineer magnetic frustration within arrays [205], and excite coupled arrays with
spin transfer torques arising from spin-polarised currents injected from a magnetic
STM tip [206]. Quantum tunnelling between Néel states of antiferromagnetic chains
has been observed [207]. Building a chain of Fe atoms on a superconducting Pb
surface has allowed for the detection of Majorana fermion-like collective excitations
in a condensed matter system [208], manifesting themselves as zero-energy states
at the ends of the chain.

Anisotropy of individual atoms

The thermal stability of any magnetic moment arises from its magnetic anisotropy
energy . At the atomic scale, this comes from the ligand crystal fields (Chapter ANI),
the electrostatic energy arising from the shape of the orbitals of both the magnetic
adatom and its neighbours in the substrate. Spin-orbit coupling locks the adatom
spin moment to its orbital moment, which is subject to these ligand fields, generating
the anisotropy. From this we see that the three factors needed for a strong anisotropy
are strong spin-orbit coupling , a large orbital moment, and a strong ligand field. The
last two are usually inversely correlated to each other, since the ligand field tends to
quench the orbital moment.
Single cobalt atoms deposited onto platinum (111) were found to have a giant
magnetic anisotropy energy of 9 meV per atom, determined using x-ray magnetic
circular dichroism [209]. Here the large unquenched orbital moment of 1.1 Bohr
magnetons and strong spin-orbit coupling induced by the heavy metal substrate
combine to yield this large value. The use of STM-based methods described
above show that even such large anisotropy energies are not enough to thermally
stabilise individual atomic moments on a metallic substrate [210, 203, 202], due to
21 Artificially Engineered Magnetic Materials 1071

spin-flip scattering of electrons in the substrate. Partially isolating the the adatom
electronically from the susbtrate using interlayers such as copper nitride [211] or
magnesium oxide [212] was found to boost the anisotropy considerably: indeed in
the latter case a record anisotropy energy of 60 meV for Co, close to the free atom
value, was achieved. Tuning the strength of the interactions between adatom and
substrate allows the strength of the anisotropy, in turn, to be controlled [213].

References
1. Y. Yoshizawa, S. Oguma, K. Yamauchi, New Fe-based soft magnetic alloys composed of
ultrafine grain structure. J. Appl. Phys. 64, 6044 (1988)
2. M. Manivel Raja, K. Chattopadhyay, B. Majumdar, A. Narayanasamy, Structure and soft
magnetic properties of FINEMET alloys. J. Alloys Compounds 297, 199 (2000)
3. G. Herzer, Grain structure and magnetism of nanocrystalline ferromagnets. IEEE Trans.
Magn. 25, 3327 (1989)
4. R. Alben, J.J. Becker, M.C. Chi, Random anisotropy in amorphous ferromagnets. Journal of
Applied Physics 49, 1653 (1978)
5. W.H. Meiklejohn, C.P. Bean, New magnetic anisotropy. Phys. Rev. 105, 904 (1957)
6. J. Nogués, I.K. Schuller, Exchange bias. J. Magn. Magn. Mater. 192, 203 (1999)
7. R.L. Stamps, Mechanisms for exchange bias. J. Phys. D: Appl. Phys. 33, R247 (2000)
8. M. Kiwi, Exchange bias theory. J. Magn. Magn. Mater. 234, 584 (2001)
9. B. Dieny, V.S. Speriosu, S.S.P. Parkin, B.A. Gurney, D.R. Wilhoit, D. Mauri, Giant magne-
toresistive in soft ferromagnetic multilayers. Phys. Rev. B 43, 1297 (1991)
10. P. Steadman, M. Ali, A.T. Hindmarch, C.H. Marrows, H.B. J., S. Langridge, R.M. Dalgliesh,
S. Foster, Exchange bias in spin-engineered double superlattices. Phys. Rev. Lett. 89, 077201
(2002)
11. D. Mauri, H.C. Siegmann, P.S. Bagus, E. Kay, Simple model for thin ferromagnetic films
exchange coupled to an antiferromagnetic substrate. J. Appl. Phys. 62, 3047 (1987)
12. A.P. Malozemoff, Random-field model of exchange anisotropy at rough ferromagnetic-
antiferromagnetic interfaces. Phys. Rev. B 35, 3679 (1987)
13. E. Fulcomer, S.H. Charap, Thermal fluctuation aftereffect model for some systems with
ferromagnetic-antiferromagnetic coupling. J. Appl. Phys. 43, 4190 (1972)
14. C. Park, K. Lee, D. Hwang, S. Lee, M. Kim, J. Rhee, Exchange coupling between
antiferromagnetic NiO and ferromagnetic films. Journal of the Korean Physical Society 31,
508 (1997)
15. R.F.C. Farrow, R.F. Marks, S. Gider, A.C. Marley, S.S.P. Parkin, D. Mauri, Mnx Pt1−x : A new
exchange bias material for Permalloy. J. Appl. Phys. 81, 4986 (1997)
16. R. Jungblut, R. Coehoorn, M.T. Johnson, J. aan de Stegge, A. Reinders, Orientational
dependence of the exchange biasing in molecular-beam-epitaxy-grown Ni80 Fe20 /Fe50 Mn50
bilayers (invited). J. Appl. Phys. 75, 6659 (1994)
17. R. Nakatani, H. Hoshiya, K. Hoshino, Y. Sugita, Relationship between film structure and
exchange coupling in Mn-Ir/Ni-Fe films. J. Magn. Magn. Mater. 173, 321 (1997)
18. M. Ali, P. Adie, C.H. Marrows, D. Greig, B.J. Hickey, R.L. Stamps, Exchange bias using a
spin glass. Nat. Mater. 6, 70 (2007)
19. C. Maurain, Sur une action magnétisante de contact et son rayon d’activité. J. Phys 1, 90
(1902)
20. J.C. Bruyére, G. Clerc, O. Massenet, D. Paccard, R. Montmory, L. Néel, J. Valin, A. Yelon,
Indirectly coupled films. IEEE Trans. Magn. 1, 174 (1965)
21. O. Massenet, F. Biraget, H. Juretschke, R. Montmory., A. Yelon, Orgin of coupling in
multilayered films. IEEE Trans. Magn. 2, 553 (1966)
1072 C. H. Marrows

22. S.S.P. Parkin, Systematic variation of the strength and oscillation period of indirect magnetic
exchange coupling through the 3d, 4d, and 5d transition metals. Phys. Rev. Lett. 67, 3598
(1991)
23. P. Grünberg, R. Schreiber, Y. Pang, M.B. Brodsky, H. Sowers, Layered magnetic structures:
Evidence for antiferromagnetic coupling of Fe layers across Cr interlayers. Phys. Rev. Lett.
57, 2442 (1986)
24. S.S.P. Parkin, N. More, K.P. Roche, Oscillations in exchange coupling and magnetoresistance
in metallic superlattice structures: Co/Ru, Co/Cr, and Fe/Cr. Phys. Rev. Lett. 64, 2304 (1990)
25. C. Kittel, Indirect exchange interactions in metals. Solid State Phys. 22, 1 (1968)
26. R. Coehoorn, Period of oscillatory exchange interactions in Co/Cu and Fe/Cu multilayer
systems. Phys. Rev. B 44, 9331 (1991)
27. Y. Yafet, Ruderman-Kittel-Kasuya-Yosida range function of a one-dimensional free-electron
gas. Phys. Rev. B 36, 3948 (1987)
28. P. Bruno, C. Chappert, Ruderman-Kittel theory of oscillatory interlayer exchange coupling.
Phys. Rev. B 46, 261 (1992)
29. D.M. Edwards, J. Mathon, R.B. Muniz, M.S. Phan, Oscillations of the exchange in magnetic
multilayers as an analog of de Haas-van Alphen effect. Phys. Rev. Lett. 67, 493 (1991)
30. J.E. Ortega, F.J. Himpsel, Quantum well states as mediators of magnetic coupling in
superlattices. Phys. Rev. Lett. 69, 844 (1992)
31. P. Bruno, Interlayer exchange coupling: a unified physical picture. J. Magn. Magn. Mater.
121, 248 (1993)
32. K.B. Hathaway, J.R. Cullen, A free-electron model for the exchange coupling of ferromagnets
through paramagnetic metals. J. Magn. Magn. Mater. 104-107, 1840 (1992)
33. M. Pinarbasi, H.A.A. Santini, E.E. Marinero, P.C. Arnett, R.W. Olson, R. Hsiao, M.L.
Williams, R.N. Payne, R.H.H. Wang, J.O. Moore, B.A. Gurney, T. Lin, R.E.E. Fontana, 12
Gb/in2 recording demonstration with SV read heads and conventional narrow pole-tip write
heads. IEEE Trans. Magn. 35, 689 (1999)
34. J. Slaughter, R. Dave, M. Durlam, G. Kerszykowski, K. Smith, K. Nagel, B. Feil, J. Calder, M.
DeHerrera, B. Garni, S. Tehrani, in Electron Devices Meeting, 2005. IEDM Technical Digest.
IEEE International (2005), p. 873. doi: https://doi.org/10.1109/IEDM.2005.1609496
35. S.H. Yang, K.S. Ryu, S.S.P. Parkin, Domain-wall velocities of up to 750 ms−1 driven by
exchange-coupling torque in synthetic antiferromagnets. Nature Nanotech. 10, 221 (2015)
36. L. Néel, L’anisotropie superficielle des substances ferromagnétiques. Compt. Rend. 237, 1468
(1953)
37. P.F. Carcia, A.D. Meinhaldt, A. Suna, Perpendicular magnetic anisotropy in Pd/Co thin film
layered structures. Appl. Phys. Lett. 47, 178 (1985)
38. W.J.M. de Jonge, P.J.H. Bloemen, F.J.A. den Broeder, Experimental Investigations of
Magnetic Anisotropy (Springer, Berlin, 1994), Ultrathin Magnetic Structures, vol. 1, chap.
2.3, p. 65
39. G.H.O. Daalderop, P.J. Kelly, F.J.A. den Broeder, Prediction and confirmation of perpendicu-
lar magnetic anisotropy in Co/Ni multilayers. Phys. Rev. Lett. 68, 682 (1992)
40. S. Ikeda, K. Miura, H. Yamamoto, K. Mizunuma, H.D. Gan, M. Endo, S. Kanai, J. Hayakawa,
F.M. andH. Ohno, A perpendicular-anisotropy CoFeB-MgO magnetic tunnel junction. Nature
Mater. 9, 721 (2010)
41. I. Dzyaloshinsky, A thermodynamic theory of “weak” ferromagnetism of antiferromagnetics.
J. Phys. Chem. Solids 4, 241 (1958)
42. T. Moriya, New mechanism of anisotropic superexchange interaction. Phys. Rev. Lett. 4, 228
(1960)
43. T. Moriya, Anisotropic superexchange interaction and weak ferromagnetism. Phys. Rev. 120,
91 (1960)
44. A. Fert, Magnetic and transport properties of metallic multilayers. Mater. Sci. Forum 59, 439
(1991)
45. A. Crépieux, C. Lacroix, Dzyaloshinsky-Moriya interactions induced by symmetry breaking
at a surface. J. Magn. Magn. Mater. 182, 341 (1998)
21 Artificially Engineered Magnetic Materials 1073

46. A. Fert, V. Cros, J. Sampaio, Skyrmions on the track. Nat. Nanotech. 8, 152 (2013)
47. M. Bode, M. Heide, K. von Bergmann, P. Ferriani, S. Heinze, G. Bihlmayer, A. Kubetzka, O.
Pietzsch, S. Blügel, R. Wiesendanger, Chiral magnetic order at surfaces driven by inversion
asymmetry. Nature 447, 190 (2007)
48. A. Thiaville, S. Rohart, Émilie Jué, V. Cros, A. Fert, Dynamics of Dzyaloshinskii domain
walls in ultrathin magnetic films. EPL (Europhysics Letters) 100, 57002 (2012)
49. A. Kubetzka, O. Pietzsch, M. Bode, R. Wiesendanger, Spin-polarized scanning tunneling
microscopy study of 360◦ walls in an external magnetic field. Phys. Rev. B 67, 020401
(2003)
50. G. Chen, T. Ma, A.T. N’Diaye, H. Kwon, C. Won, Y. Wu, A.K. Schmid, Tailoring the chirality
of magnetic domain walls by interface engineering. Nat. Commun. 4, 2671 (2013)
51. M.J. Benitez, A. Hrabec, A.P. Mihai, T.A. Moore, G. Burnell, D. McGrouther, C.H. Marrows,
S. McVitie, Magnetic microscopy and topological stability of homochiral Néel domain walls
in a Pt/Co/AlOx trilayer. Nat. Commun. 6, 8957 (2015)
52. A. Bogdanov, A. Hubert, Thermodynamically stable magnetic vortex states in magnetic
crystals. J. Magn. Magn. Mater. 138, 255 (1994)
53. R. Wiesendanger, Nanoscale magnetic skyrmions in metallic films and multilayers: a new
twist for spintronics. Nat. Reviews Mater. 1, 16044 (2016)
54. S. Heinze, K. von Bergmann, M. Menzel, J. Brede, A. Kubetzka, R. Wiesendanger, G.
Bihlmayer, S. Blügel, Spontaneous atomic-scale magnetic skyrmion lattice in two dimen-
sions. Nat. Phys. 7, 713 (2011)
55. G. Chen, A. Mascaraque, A.T. N’Diaye, A.K. Schmid, Room temperature skyrmion ground
state stabilized through interlayer exchange coupling. Appl. Phys. Lett. 106, 242404
(2015)
56. C. Moreau-Luchaire, C. Moutafis, N. Reyren, J. Sampaio, C.A.F. Vaz, N.V. Horne, K.
Bouzehouane, K. Garcia, C. Deranlot, P. Warnicke, P. Wohlhüter, J.M. George, M. Weigand,
J. Raabe, V. Cros, A. Fert, Additive interfacial chiral interaction in multilayers for stabilization
of small individual skyrmions at room temperature. Nat. Nanotech. 11, 444 (2016)
57. O. Boulle, J. Vogel, H. Yang, S. Pizzini, D. de Souza Chaves, A. Locatelli, T.O. Menteş,
A. Sala, L.D. Buda-Prejbeanu, O. Klein, M. Belmeguenai, Y. Roussigné, A. Stashkevich,
S.M. Chérif, L. Aballe, M. Foerster, M. Chshiev, S. Auffret, I.M. Miron, G. Gaudin, Room-
temperature chiral magnetic skyrmions in ultrathin magnetic nanostructures. Nat. Nanotech.
11, 449 (2016)
58. W. Jiang, P. Upadhyaya, W. Zhang, G. Yu, M.B. Jungfleisch, F.Y. Fradin, J.E. Pearson, Y.
Tserkovnyak, K.L. Wang, O. Heinonen, S.G.E. te Velthuis, A. Hoffmann, Blowing magnetic
skyrmion bubbles. Science 349, 283 (2015)
59. S. Woo, K. Litzius, B. Krüger, M.Y. Im, L. Caretta, K. Richter, M. Mann, A. Krone,
R.M. Reeve, M. Weigand, P. Agrawal, I. Lemesh, M.A. Mawass, P. Fischer, M. Kläui,
G.S.D. Beach, Observation of room-temperature magnetic skyrmions and their current-driven
dynamics in ultrathin metallic ferromagnets. Nat. Mater. 15, 501 (2016)
60. W. Shockley, G.L. Pearson, Modulation of conductance of thin films of semi-conductors by
surface charges. Phys. Rev. 74, 232 (1948)
61. K. von Klitzing, G. Dorda, M. Pepper, New method for high-accuracy determination of the
fine-structure constant based on quantized hall resistance. Phys. Rev. Lett. 45, 494 (1980)
62. D.C. Tsui, H.L. Stormer, A.C. Gossard, Two-dimensional magnetotransport in the extreme
quantum limit. Phys. Rev. Lett. 48, 1559 (1982)
63. K.S. Novoselov, Z. Jiang, Y. Zhang, S.V. Morozov, H.L. Stormer, U. Zeitler, J.C. Maan, G.S.
Boebinger, P. Kim, A.K. Geim, Room-temperature quantum hall effect in graphene. Science
315, 1379 (2007)
64. A. Ohtomo, H.Y. Hwang, A high-mobility electron gas at the LaAlO3 /SrTiO3 heterointerface.
Nature 427, 423 (2004)
65. A.D. Caviglia, S. Gariglio, C. Cancellieri, B. Sacépé, A. Fête, N. Reyren, M. Gabay, A.F.
Morpurgo, J.M. Triscone, Two-dimensional quantum oscillations of the conductance at
laalo3 /srtio3 interfaces. Phys. Rev. Lett. 105, 236802 (2010)
1074 C. H. Marrows

66. N. Reyren, S. Thiel, A.D. Caviglia, L.F. Kourkoutis, G. Hammerl, C. Richter, C.W. Schneider,
T. Kopp, A.S. Rüetschi, D. Jaccard, M. Gabay, D.A. Muller, J.M. Triscone, J. Mannhart,
Superconducting interfaces between insulating oxides. Science 317, 1196 (2007)
67. S. Thiel, G. Hammerl, A. Schmehl, C.W. Schneider, J. Mannhart, Tunable quasi-two-
dimensional electron gases in oxide heterostructures. Science 313, 1942 (2006)
68. A.D. Caviglia, S. Gariglio, N. Reyren, D. Jaccard, T. Schneider, M. Gabay, S. Thiel, G.
Hammerl, J. Mannhart, J.M. Triscone, Electric field control of the LaAlO3 /SrTiO3 interface
ground state. Nature 456, 624 (2008)
69. A. Brinkman, M. Huijben, M. van Zalk, J. Huijben, U. Zeitler, J.C. Maan, W.G. van der Wiel,
G. Rijnders, D.H.A. Blank, H. Hilgenkamp, Magnetic effects at the interface between non-
magnetic oxides. Nat. Mater. 6, 493 (2007)
70. S. Seri, L. Klein, Antisymmetric magnetoresistance of the SrTiO3 /LaAlO3 interface. Phys.
Rev. B 80, 180410 (2009)
71. Ariando, X. Wang, G. Baskaran, Z.Q. Liu, J. Huijben, J.B. Yi, A. Annadi, A.R. Barman,
A. Rusydi, S. Dhar, Y.P. Feng, J. Ding, H. Hilgenkamp, T. Venkatesan, Electronic phase
separation at the LaAlO3 /SrTiO3 interface. Nat. Commun. 2, 188 (2011)
72. J.A. Bert, B. Kalisky, C. Bell, M. Kim, Y. Hikita, H.Y. Hwang, K.A. Moler, Direct imaging
of the coexistence of ferromagnetism and superconductivity at the LaAlO3 /SrTiO3 interface.
Nat. Phys. 7, 767 (2011)
73. D.A. Dikin, M. Mehta, C.W. Bark, C.M. Folkman, C.B. Eom, V. Chandrasekhar, Coexistence
of superconductivity and ferromagnetism in two dimensions. Phys. Rev. Lett. 107, 056802
(2011)
74. B. Kalisky, J.A. Bert, B.B. Klopfer, C. Bell, H.K. Sato, M. Hosoda, Y. Hikita, H.Y. Hwang,
K.A. Moler, Critical thickness for ferromagnetism in LaAlO3 /SrTiO3 heterostructures. Nat.
Commun. 3, 922 (2012)
75. J.E. Moore, The birth of topological insulators. Nature 464, 194 (2010)
76. C.L. Kane, E.J. Mele, Z2 topological order and the quantum spin Hall effect. Phys. Rev. Lett.
95, 146802 (2005)
77. L. Fu, C.L. Kane, Topological insulators with inversion symmetry. Phys. Rev. B 76, 045302
(2007)
78. D. Hsieh, D. Qian, L. Wray, Y. Xia, Y.S. Hor, R.J. Cava, M.Z. Hasan, A topological dirac
insulator in a quantum spin hall phase. Nature 452, 970 (2008)
79. C.H. Li, O.M.J. van ’t Erve, J.T. Robinson, Y. Liu, L. Li, B.T. Jonker, Electrical detection of
charge-current-induced spin polarization due to spin-momentum locking in Bi2 Se3 . Nature
Nanotech. 9, 218 (2014)
80. Experiments provide first direct signatures of a topological insulator - a new phase of quantum
matter (2009). URL http://www-ssrl.slac.stanford.edu/content/science/highlight/2009-03-30/
experiments-provide-first-direct-signatures-topological-insulator-new. SSRL Science High-
light
81. R. Yu, W. Zhang, H.J. Zhang, S.C. Zhang, X. Dai, Z. Fang, Quantized anomalous hall effect
in magnetic topological insulators. Science 329, 61 (2010)
82. C.Z. Chang, J. Zhang, X. Feng, J. Shen, Z. Zhang, M. Guo, K. Li, Y. Ou, P. Wei, L.L. Wang,
Z.Q. Ji, Y. Feng, S. Ji, X. Chen, J. Jia, X. Dai, Z. Fang, S.C. Zhang, K. He, Y. Wang, L. Lu,
X.C. Ma, Q.K. Xue, Experimental observation of the quantum anomalous Hall effect in a
magnetic topological insulator. Science 340, 167 (2013)
83. D.A. Allwood, G. Xiong, C.C. Faulkner, D. Atkinson, D. Petit, R.P. Cowburn, Magnetic
domain-wall logic. Science 309, 1688 (2005)
84. S.S.P. Parkin, M. Hayashi, L. Thomas, Magnetic domain-wall racetrack memory. Science
320, 190 (2008)
85. A. Fert, L. Piraux, Magnetic nanowires. J. Magn. Magn. Mater. 200, 338 (1999)
86. E. Ferain, R. Legras, Track-etched membrane: dynamics of pore formation. Nucl. Instrum.
Methods B 84, 331 (1994)
87. H. Masuda, K. Fukuda, Ordered metal nanohole arrays made by a two-step replication of
honeycomb structures of anodic alumina. Science 268, 1466 (1995)
21 Artificially Engineered Magnetic Materials 1075

88. K. Nielsch, F. Müller, A.P. Li, U. Gösele, Uniform nickel deposition into ordered alumina
pores by pulsed electrodeposition. Adv. Mater. 12, 582 (2000)
89. A. Blondel, B. Doudin, J.P. Ansermet, Comparative study of the magnetoresistance of
electrodeposited Co/Cu multilayered nanowires made by single and dual bath techniques.
J. Magn. Magn. Mater. 165, 34 (1997)
90. M. Alper, K. Attenborough, R. Hart, S.J. Lane, D.S. Lashmore, C. Younes, W. Schwarzacher,
Giant magnetoresistance in electrodeposited superlattices. Appl. Phys. Lett. 63, 2144 (1993)
91. D. AlMawlawi, N. Coombs, M. Moskovits, Magnetic properties of Fe deposited into anodic
aluminum oxide pores as a function of particle size. J. Appl. Phys. 70, 4421 (1991)
92. J.C. Lodder, L. Cheng-Zhang, Reversal behaviour in perpendicular iron particle arrays
(alumite media). IEEE Trans. Magn. 25, 4171 (2002)
93. T.M. Whitney, J.S. Jiang, P.C. Searson, C.L. Chien, Fabrication and magnetic properties of
arrays of metallic nanowires. Science 261, 1316 (1993)
94. L. Piraux, S. Dubois, E. Ferain, R. Legras, K. Ounadjela, J.M. George, J.L. Maurice, A. Fert,
Anisotropic transport and magnetic properties of arrays of sub-micron wires. J. Magn. Magn.
Mater. 165, 352 (1997)
95. L. Piraux, J.M. George, J.F. Despres, C. Leroy, E. Ferain, R. Legras, K. Ounadjela, A. Fert,
Giant magnetoresistance in magnetic multilayered nanowires. Appl. Phys. Lett. 65, 2484
(1994)
96. L. Piraux, S. Dubois, A. Fert, Perpendicular giant magnetoresistance in magnetic multilayered
nanowires. J. Magn. Magn. Mater. 159, L287 (1996)
97. L. Piraux, S. Dubois, A. Fert, L. Belliard, The temperature dependence of the perpendicular
giant magnetoresistance in Co/Cu multilayered nanowires. Eur. Phys. J. B 4, 413 (1998)
98. J.E. Wegrowe, D. Kelly, X. Hoffer, P. Guittienne, J.P. Ansermet, Tailoring anisotropic
magnetoresistance and giant magnetoresistance hysteresis loops with spin-polarized current
injection. J. Appl. Phys. 89, 7127 (2001)
99. X. Huang, L. Tan, H. Cho, B.J.H. Stadler, Magnetoresistance and spin transfer torque in
electrodeposited Co/Cu multilayered nanowire arrays with small diameters. J. Appl. Phys.
105, 07D128 (2009)
100. C.L. Kane, E.J. Mele, Quantum spin Hall effect in graphene,. Phys. Rev. Lett. 95, 226801
(2005)
101. B.A. Bernevig, S.C. Zhang, Quantum spin Hall effect. Phys. Rev. Lett. 96, 106802 (2006)
102. M. König, S. Wiedmann, C. Brüne, A. Roth, H. Buhmann, L.W. Molenkamp, X.L. Qi, S.C.
Zhang, Quantum spin Hall insulator state in HgTe quantum wells. Science 318, 766 (2007)
103. K. Suzuki, Y. Harada, K. Onomitsu, K. Muraki, Edge channel transport in the InAs/GaSb
topological insulating phase. Phys. Rev. B 87, 235311 (2013)
104. L. Du, I. Knez, G. Sullivan, R.R. Du, Robust helical edge transport in gated InAs/GaSb
bilayers. Phys. Rev. Lett. 114, 096802 (2015)
105. R.P. Cowburn, D.K. Koltsov, A.O. Adeyeye, M.E. Welland, D.M. Tricker, Single-domain
circular nanomagnets. Phys. Rev. Lett. 83, 1042 (1999)
106. R.P. Cowburn, A.O. Adeyeye, M.E. Welland, Controlling magnetic ordering in coupled
nanomagnet arrays. New J. Phys. 1, 16 (1999)
107. R.P. Cowburn, M.E. Welland, Room temperature magnetic quantum cellular automata.
Science 287, 1466 (2000)
108. J. Shibata, Y. Otani, Magnetic vortex dynamics in a two-dimensional square lattice of
ferromagnetic nanodisks. Phys. Rev. B 70, 012404 (2004)
109. C. Behncke, M. Hänze, C.F. Adolff, M. Weigand, G. Meier, Band structure engineering of
two-dimensional magnonic vortex crystals. Phys. Rev. B 91, 224417 (2015)
110. S.T. Bramwell, M.J.P. Gingras, Spin ice state in frustrated magnetic pyrochlore materials.
Science 294, 1495 (2001)
111. J.D. Bernal, R.H. Fowler, A theory of water and ionic solution, with particular reference to
hydrogen and hydroxyl ions. J. Chem. Phys. 1, 515 (1933)
112. L. Pauling, The structure and entropy of ice and of other crystals with some randomness of
atomic arrangement. J. Am. Chem. Soc. 57, 2680 (1935)
1076 C. H. Marrows

113. C. Castelnovo, R. Moessner, S.L. Sondhi, Magnetic monopoles in spin ice. Nature 451, 42
(2008)
114. M. Tanaka, E. Saitoh, H. Miyajima, T. Yamaoka, Y. Iye, Domain structure and magnetic ice-
order in NiFe nano-network with honeycomb structure. J. Appl. Phys. 97, 10J710 (2005)
115. R.F. Wang, C. Nisoli, R.S. Freitas, J. Li, W. McConville, B.J. Cooley, M.S. Lund, N. Samarth,
C. Leighton, V.H. Crespi, P. Schiffer, Artificial ‘spin ice’ in a geometrically frustrated lattice
of nanoscale ferromagnetic islands. Nature 439, 303 (2006)
116. Y. Han, Y. Shokef, A.M. Alsayed, P. Yunker, T.C. Lubensky, A.G. Yodh, Geometric frustration
in buckled colloidal monolayers. Nature 456, 898 (2008)
117. M.L. Latimer, G.R. Berdiyorov, Z.L. Xiao, F.M. Peeters, W.K. Kwok, Realization of artificial
ice systems for magnetic vortices in a superconducting MoGe thin film with patterned
nanostructures. Phys. Rev. Lett. 111, 067001 (2013)
118. A. Ortiz-Ambriz, P. Tierno, Engineering of frustration in colloidal artificial ices realized on
microfeatured grooved lattices. Nature Commun. 7, 10575 (2016)
119. C. Nisoli, R. Moessner, P. Schiffer, Colloquium: Artificial spin ice: Designing and imaging
magnetic frustration. Rev. Mod. Phys. 85, 1473 (2013)
120. C. Nisoli, R. Wang, J. Li, W.F. McConville, P.E. Lammert, P. Schiffer, V.H. Crespi, Ground
state lost but degeneracy found: the effective thermodynamics of ‘artificial spin ice’. Phys.
Rev. Lett. 98, 217103 (2007)
121. X. Ke, J. Li, C. Nisoli, P.E. Lammert, W. McConville, R.F. Wang, V.H. Crespi, P. Schiffer,
Energy minimization and ac demagnetization in a nanomagnet array. Phys. Rev. Lett. 101,
037205 (2008)
122. C. Nisoli, J. Li, X. Ke, D. Garand, P. Schiffer, V.H. Crespi, Effective temperature in an
interacting vertex system: Theory and experiment on artificial spin ice. Phys. Rev. Lett. 105,
047205 (2010)
123. Y. Qi, T. Brintlinger, J. Cumings, Direct observation of the ice rule in an artificial kagome
spin ice. Phys. Rev. B 77, 094418 (2008)
124. S. Ladak, D.E. Read, G.K. Perkins, L.F. Cohen, W.R. Branford, Direct observation of
magnetic monopole defects in an artificial spin-ice system. Nature Phys. 6, 359 (2010)
125. E. Mengotti, L.J. Heyderman, A. Fraile Rodríguez, F. Nolting, R.V. Hügli, H.B. Braun, Real-
space observation of emergent magnetic monopoles and associated Dirac strings in artificial
kagome spin ice. Nature Phys. 7, 68 (2011)
126. J.P. Morgan, A. Stein, S. Langridge, C.H. Marrows, Thermal ground-state ordering and
elementary excitations in artificial magnetic square ice. Nature Phys. 7, 75 (2011)
127. J.M. Porro, A.Bedoya-Pinto, A. Berger, P. Vavassori, Exploring thermally induced states in
square artificial spin-ice arrays. New J. Phys. 15, 055012 (2013)
128. S. Zhang, I. Gilbert, C. Nisoli, G.W. Chern, M.J. Erickson, L. O’Brien, C. Leighton, P.E.
Lammert, V.H. Crespi, P. Schiffer, Crystallites of magnetic charges in artificial spin ice.
Nature 500, 553 (2013)
129. Y. Perrin, B. Canals, N. Rougemaille, Extensive degeneracy, Coulomb phase and magnetic
monopoles in artificial square ice. Nature 540, 410 (2016)
130. G. Möller, R. Moessner, Magnetic multipole analysis of kagome and artificial ice dipolar
arrays. Phys. Rev. B 80, 140409 (2009)
131. G.W. Chern, P. Mellado, O. Tchernyshyov, Two-stage ordering of spins in dipolar spin ice on
the kagome lattice. Phys. Rev. Lett. 106, 207202 (2011)
132. L. Anghinolfi, H. Luetkens, J. Perron, M.G. Flokstra, O. Sendetskyi, A. Suter, T. Prokscha,
P.M. Derlet, S.L. Lee, L.J. Heyderman, Thermodynamic phase transitions in a frustrated
magnetic metamaterial. Nature Commun. 6, 8278 (2015)
133. I. Gilbert, C. Nisoli, G.W. Chern, S. Zhang, L. O’Brien, B. Fore, P. Schiffer, Emergent ice
rule and magnetic charge screening from vertex frustration in artificial spin ice. Nature Phys.
10, 670 (2014)
134. R.F. Wang, C. Nisoli, R.S. Freitas, J. Li, W. McConville, B.J. Cooley, M.S. Lund, N. Samarth,
C. Leighton, V.H. Crespi, P. Schiffer, Demagnetization protocols for frustrated interacting
nanomagnet arrays. J. Appl. Phys. 101, 09J104 (2007)
21 Artificially Engineered Magnetic Materials 1077

135. J.P. Morgan, A. Bellew, A. Stein, S. Langridge, C.H. Marrows, Linear field demagnetization
of artificial magnetic square ice. Frontiers in Physics 1, 28 (2013)
136. C. Nisoli, J. Li, X. Ke, D. Garand, P. Schiffer, V.H. Crespi, Effective temperature in an
interacting vertex system: Theory and experiment on artificial spin ice. Phys. Rev. Lett. 105,
047205 (2010)
137. J. Morgan, J. Akerman, A. Stein, C. Phatak, R.M.L. Evans, S. Langridge, C. Marrows, Real
and effective thermal equilibrium in artificial square spin ices. Phys. Rev. B 87, 024405
(2013)
138. P.E. Lammert, X. Ke, J. Li, C.Nisoli, D.M. Garand, V.H. Crespi, P. Schiffer, Direct entropy
determination and application to artificial spin ice. Nature Phys. 6, 786 (2010)
139. V. Kapaklis, U.B. Arnalds, A. Harman-Clarke, E.T. Papaioannou, M. Karimipour, P. Korelis,
A. Taroni, P.C.W. Holdsworth, S.T. Bramwell, B. Hjörvarsson, Melting artificial spin ice. New
J. Phys. 14 (2012)
140. I.A. Chioar, B. Canals, D. Lacour, M. Hehn, B. Santos Burgos, T.O. Menteş, A. Locatelli,
F. Montaigne, N. Rougemaille, Kinetic pathways to the magnetic charge crystal in artificial
dipolar spin ice. Phys. Rev. B 90, 220407 (2014)
141. J. Drisko, S. Daunheimer, J. Cumings, FePd3 as a material for studying thermally active
artificial spin ice systems. Phys. Rev. B 91, 224406 (2015)
142. A. Farhan, P.M. Derlet, A. Kleibert, A. Balan, R.V. Chopdekar, M. Wyss, L. Anghinolfi, F.
Nolting, L.J. Heyderman, Exploring hyper-cubic energy landscapes in thermally active finite
spin-ice systems. Nature Phys. 9, 375 (2013)
143. A. Farhan, P.M. Derlet, A. Kleibert, A. Balan, R.V. Chopdekar, M. Wyss, J. Perron, A. Scholl,
F. Nolting, L.J. Heyderman, Direct observation of thermal relaxation in artificial spin ice.
Phys. Rev. Lett. 111, 057204 (2013)
144. V. Kapaklis, U.B. Arnalds, A. Farhan, R.V.C.A. Balan, A. Scholl, L.J. Heyderman, B.
Hjörvarsson, Thermal fluctuations in artificial spin ice. Nature Nano. 9, 514 (2014)
145. S.A. Morley, A. Stein, M.C. Rosamond, D. Alba Venero, A. Hrabec, P.M. Shepley, M.Y.
Im, P. Fischer, M.T. Bryan, D.A. Allwood, P. Steadman, S. Langridge, C.H. Marrows,
Temperature and magnetic-field driven dynamics in artificial magnetic square ice. Proc. SPIE
9551, 95511Q (2015)
146. M.J. Morrison, T.R. Nelson, C. Nisoli, Unhappy vertices in artificial spin ice: new degenera-
cies from vertex frustration. New J. Phys. 15, 045009 (2013)
147. I. Gilbert, Y. Lao, I. Carrasquillo, L. O’Brien, J.D. Watts, M. Manno, C. Leighton, A. Scholl,
C. Nisoli, P. Schiffer, Emergent reduced dimensionality by vertex frustration in artificial spin
ice. Nature Phys. 12, 162 (2016)
148. B. Canals, I.A. Chioar, V.D. Nguyen, M. Hehn, D. Lacour, F. Montaigne, A. Locatelli, T.O.
Menteş, B.S. Burgos, N. Rougemaille, Fragmentation of magnetism in artificial kagome
dipolar spin ice. Nature Commun. 7, 11446 (2016)
149. B. Farmer, V.S. Bhat, A. Balk, E. Teipel, N. Smith, J. Unguris, D.J. Keavney, J.T. Hastings,
L.E.D. Long, Direct imaging of coexisting ordered and frustrated sublattices in artificial
ferromagnetic quasicrystals. Phys. Rev. B 93, 134428 (2016)
150. Y.L. Wang, Z.L. Xiao, A. Snezhko, J. Xu, L.E. Ocola, R. Divan, J.E. Pearson, G.W. Crabtree,
W.K. Kwok, Rewritable artificial magnetic charge ice. Science 352, 962 (2016)
151. M.A. Kastner, Artificial atoms. Physics Today 46, 24 (1993)
152. R.C. Ashoori, Electrons in artificial atoms. Nature 379, 413 (1996)
153. K.K. Likharev, Single-electron devices and their applications. Proc. IEEE 87, 606 (1999)
154. P. Seneor, A. Bernand-Mantel, F. Petroff, Nanospintronics: when spintronics meets single
electron physics. J. Phys.: Cond. Matt. 19, 165222 (2007)
155. K.J. Dempsey, D. Ciudad, C.H. Marrows, Single electron spintronics. Philosophical Trans-
actions of the Royal Society A: Mathematical, Physical and Engineering Sciences 369, 3150
(2011)
156. D.V. Averin, Y.V. Nazarov, in Single Charge Tunneling: Coulomb Blockade Phenomena in
Nanostructures, NATO ASI Series B: Physics, vol. 294, ed. by H. Grabert, M.H. Devoret
(Plenum Press, New York, 1992), chap. 6, pp. 217–247
1078 C. H. Marrows

157. K.K. Likharev, Correlated discrete charge transfer of single electrons in ultrasmall tunnel
junctions. IBM. J. Res. Dev. 32, 144 (1988)
158. D.V. Averin, Y.V. Nazarov, Virtual electron diffusion during quantum tunneling of the electric
charge. Phys. Rev. Lett. 65, 2446 (1990)
159. J. Barnaś, I. Weymann, Spin effects in single-electron tunnelling. J. Phys.: Cond. Matt. 20,
423202 (2008)
160. J. Barnaś, A. Fert, Magnetoresistance oscillations due to charging effects in double ferromag-
netic tunnel junctions. Phys. Rev. Lett. 80, 1058 (1998)
161. J. Barnaś, A. Fert, Effects of spin accumulation on single-electron tunneling in a double
ferromagnetic microjunction. Europhys. Lett. 44, 85 (1998)
162. M. Jullière, Tunneling between ferromagnetic films. Phys. Lett. A 54, 225 (1975)
163. K. Yakushiji, F. Ernult, H. Imamura, K. Yamane, S. Mitani, K. Takanashi, S. Takahashi,
S. Maekawa, H. Fujimori, Enhanced spin accumulation and novel magnetotransport in
nanoparticles. Nature Materials 4, pp57 (2005)
164. O.D. Miller, B. Muralidharan, N. Kapur, A.W. Ghosh, Rectification by charging: Contact-
induced current asymmetry in molecular conductors. Phys. Rev. B 77, 125427 (2008)
165. A. Bernand-Mantel, P. Seneor, K. Bouzehouane, S. Fusil, C. Deranlot, F. Petroff, A. Fert,
Anisotropic magneto-Coulomb effects and magnetic single-electron-transistor action in a
single nanoparticle. Nature Physics 5, 920 (2009)
166. A. Brataas, Y.V. Nazarov, J. Inoue, G.E.W. Bauer, Spin accumulation in small ferromagnetic
double-barrier junctions. Phys. Rev. B 59, 93 (1999)
167. J. Barnaś, A. Fert, Interplay of spin accumulation and Coulomb blockade in double
ferromagnetic junctions. J. Magn. Magn. Mater. 192, 391 (1999)
168. H. Imamura, S. Takahashi, S. Maekawa, Spin-dependent Coulomb blockade in
ferromagnet/normal-metal/ferromagnet double tunnel junctions. Phys. Rev. B 59, 6017 (1999)
169. R.C. Temple, C.H. Marrows, Single-electron spin interplay for characterization of magnetic
double tunnel junctions. Physical Review B 88 (2013)
170. R.C. Temple, M. McLaren, R.M.D. Brydson, B.J. Hickey, C.H. Marrows, Long spin lifetime
and large barrier polarisation in single electron transport through a CoFe nanoparticle.
Scientific Reports 6, 28296 (2016)
171. F. Ernult, K. Yamane, S. Mitani, K. Yakushiji, K. Takanashi, Y.K. Takahashi, K. Hono, Spin-
dependent single-electron-tunneling effects in epitaxial Fe nanoparticles. Appl. Phys. Lett.
84, 3106 (2004)
172. A. Bernand-Mantel, P. Seneor, N. Lidgi, M.M. noz, V. Cros, S. Fusil, K. Bouzehouane, C.
Deranlot, A. Vaures, F. Petroff, A. Fert, Evidence for spin injection in a single metallic
nanoparticle: A step towards nanospintronics. Appl. Phys. Lett. 89, 062502 (2006)
173. Y.G. Wei, C.E. Malec, D. Davidović, Saturation of spin-polarized current in nanometer scale
aluminum grains. Phys. Rev. B 76, 195327 (2007)
174. Y. Yafet, in Solid State Physics: Advances in Research and Applications, vol. 14, ed. by F.
Seitz, D. Turnbull (Academic Press, New York, 1963), chap. 1, pp. 1–98
175. S. Mitani, Y. Nogi, H. Wang, K. Yakushiji, F. Ernult, K. Takanashi, Current-induced tunnel
magnetoresistance due to spin accumulation in Au nanoparticles. Appl. Phys. Lett. 92, 152509
(2008)
176. L.J. Geerligs, D.V. Averin, J.E. Mooij, Observation of macroscopic quantum tunneling
through the Coulomb energy barrier. Phys. Rev. Lett. 65, 3037 (1990)
177. H. Sukegawa, S. Nakamura, A. Hirohata, N. Tezuka, K. Inomata, Significant magnetore-
sistance enhancement due to a cotunneling process in double tunnel junction with single
discontinuous ferromagnetic layer insertion. Phys. Rev. Lett. 94, 068304 (2005)
178. K.J. Dempsey, A.T. Hindmarch, H.X. Wei, Q.H. Qin, Z.C. Wen, W.X. Wang, G. Vallejo-
Fernandez, D.A. Arena, X.F. Han, C.H. Marrows, Cotunneling enhancement of magne-
toresistance in double magnetic tunnel junctions with embedded superparamagnetic nife
nanoparticles. Phys. Rev. B 82, 214415 (2010)
179. C.M. Hurd, Eletrons in Metals (Wiley, New York, 1975)
180. J. Kondo, Resistance minimum in dilute magnetic alloys. Prog. Theor. Phys. 32, 37 (1964)
21 Artificially Engineered Magnetic Materials 1079

181. D. Goldhaber-Gordon, H. Shtrikman, D. Mahalu, D. Abusch-Magder, U. Meirav, M.A.


Kastner, Kondo effect in a single-electron transistor. Nature 391, 156 (1998)
182. J. Kohlhepp, P. LeClair, H. Swagten, W. de Jonge, Interfacial sensitivity and zero-bias
anomalies in magnetic tunnel junctions. Phys. Stat. Sol. (a) 189, 585 (2002)
183. S. Bermon, C.K. So, Conductance peaks produced by Kondo scattering from O2 in M-I-M
tunnel junctions. Solid State Comms. 27, 723 (1978)
184. S.Y. Bae, S.X. Wang, Transport in magnetically doped magnetic tunnel juctions. IEEE Trans.
Magn. 38, 2721 (2002)
185. C. Romeike, M. Wegewijs, W. Hofstetter, H. Schoeller, Kondo-transport spectroscopy of
single molecule magnets. Phys. Rev. Lett. 97, 206601 (2006)
186. C. Romeike, M. Wegewijs, W. Hofstetter, H. Schoeller, Quantum-tunneling-induced Kondo
effect in single molecular magnets. Phys. Rev. Lett. 96, 196601 (2006)
187. H. Yang, S.H. Yang, S.S.P. Parkin, Crossover from Kondo-assisted suppression to co-
tunneling enhancement of tunneling magnetoresistance via ferromagnetic nanodots in MgO
tunnel barrier. Nano Lett. 8, 340 (2008)
188. M. Reyes Calvo, J. Fernández-Rossier, J.J. Palacios, D. Jacob, D. Natelson, C. Untiedt, The
Kondo effect in ferromagnetic atomic contacts. Nature 458, 1150 (2009)
189. J. Weis, in CFN Lectures on Functional Nanostructures vol. 1, Lecture Notes in Physics, vol.
658, ed. by K. Busch, A.K. Powell, C. Röthig, G. Schön, J. Weissmüller (Springer, Berlin,
2005), chap. 5, pp. 87–122
190. D. Ciudad, Z.C. Wen, A.T. Hindmarch, E. Negusse, D.A. Arena, X.F. Han, C.H. Marrows,
Competition between cotunneling, kondo effect, and direct tunneling in discontinuous high-
anisotropy magnetic tunnel junctions. Phys. Rev. B 85 (2012)
191. K. Ono, H. Shimada, Y. Ootuka, Enhanced magnetic valve effect and magneto-Coulomb
oscillations in ferromagnetic single electron transistor. J. Phys. Soc. Japan 66, 1261 (1997)
192. K. Ono, H. Shimada, Y. Ootuka, Spin polarization and magneto-Coulomb oscillations in
ferromagnetic single electron devices. J. Phys. Soc. Japan 67, 2852 (1998)
193. S.J. van der Molen, N. Tombros, B.J. van Wees, Magneto-Coulomb effect in spin-valve
devices. Phys. Rev. B 73, 220406 (2006)
194. J. Wunderlich, T. Jungwirth, B. Kaestner, A.C. Irvine, A.B. Shick, N. Stone, K.Y. Wang, U.
Rana, A.D. Giddings, C.T. Foxon, R.P. Campion, D.A. Williams, B.L. Gallagher, Coulomb
blockade anisotropic magnetoresistance effect in a (Ga, Mn) As single-electron transistor.
Phys. Rev. Lett. 97, 077201 (2006)
195. R.S. Becker, J.A. Golovchenko, B.S. Swartzentruber, Atomic-scale surface modifications
using a tunnelling microscope. Nature 325, 419 (1987)
196. D.M. Eigler, E.K. Schweizer, Positioning single atoms with a scanning tunnelling microscope.
Nature 344, 524 (1990)
197. M.F. Crommie, C.P. Lutz, D.M. Eigler, Confinement of electrons to quantum corrals on a
metal surface. Science 262, 218 (1993)
198. D. Eigler, in Nanostructures and Quantum Effects, Springer Series in Materials Science, vol.
31 (Springer, Berlin Heidelberg, 1994), pp. 311–314
199. H.C. Manoharan, C.P. Lutz, D.M. Eigler, Quantum mirages formed by coherent projection of
electronic structure. Nature 403, 512 (2000)
200. G.A. Fiete, E.J. Heller, Colloquium: Theory of quantum corrals and quantum mirages. Rev.
Mod. Phys. 75, 933 (2003)
201. F. Meier, L. Zhou, J. Wiebe, R. Wiesendanger, Revealing magnetic interactions from single-
atom magnetization curves. Science 320, 82 (2008)
202. A.A. Khajetoorians, S. Lounis, B. Chilian, A.T. Costa, L. Zhou, D.L. Mills, J. Wiebe,
R. Wiesendanger, Itinerant nature of atom-magnetization excitation by tunneling electrons.
Physical Review Letters 106 (2011)
203. S. Loth, M. Etzkorn, C.P. Lutz, D.M. Eigler, A.J. Heinrich, Measurement of fast electron spin
relaxation times with atomic resolution. Science 329, 1628 (2010)
204. A.A. Khajetoorians, J. Wiebe, B. Chilian, R. Wiesendanger, Realizing all-spin-based logic
operations atom by atom. Science 332, 1062 (2011)
1080 C. H. Marrows

205. A.A. Khajetoorians, J. Wiebe, B. Chilian, S. Lounis, S. Blügel, R. Wiesendanger, Atom-by-


atom engineering and magnetometry of tailored nanomagnets. Nature Physics 8, 497 (2012)
206. A.A. Khajetoorians, B. Baxevanis, C. Hubner, T. Schlenk, S. Krause, T.O. Wehling, S. Lounis,
A. Lichtenstein, D. Pfannkuche, J. Wiebe, R. Wiesendanger, Current-driven spin dynamics of
artificially constructed quantum magnets. Science 339, 55 (2013)
207. S. Loth, S. Baumann, C.P. Lutz, D.M. Eigler, A.J. Heinrich, Bistability in atomic-scale
antiferromagnets. Science 335, 196 (2012)
208. S. Nadj-Perge, I.K. Drozdov, J. Li, H. Chen, S. Jeon, J. Seo, A.H. MacDonald, B.A.
Bernevig, A. Yazdani, Observation of Majorana fermions in ferromagnetic atomic chains on
a superconductor. Science 346, 602 (2014)
209. P. Gambardella, S. Rusponi, M. Veronese, S.S. Dhesi, C. Grazioli, A. Dallmeyer, I. Cabria, R.
Zeller, P.H. Dederichs, K. Kern, C. Carbone, H. Brune, Giant magnetic anisotropy of single
cobalt atoms and nanoparticles. Science 300, 1130 (2003)
210. A.J. Heinrich, Single-atom spin-flip spectroscopy. Science 306, 466 (2004)
211. C.F. Hirjibehedin, C.Y. Lin, A.F. Otte, M. Ternes, C.P. Lutz, B.A. Jones, A.J. Heinrich,
Large magnetic anisotropy of a single atomic spin embedded in a surface molecular network.
Science 317, 1199 (2007)
212. I.G. Rau, S. Baumann, S. Rusponi, F. Donati, S. Stepanow, L. Gragnaniello, J. Dreiser, C.
Piamonteze, F. Nolting, S. Gangopadhyay, O.R. Albertini, R.M. Macfarlane, C.P. Lutz, B.A.
Jones, P. Gambardella, A.J. Heinrich, H. Brune, Reaching the magnetic anisotropy limit of a
3d metal atom. Science 344, 988 (2014)
213. J.C. Oberg, M.R. Calvo, F. Delgado, M. Moro-Lagares, D. Serrate, D. Jacob, J. Fernández-
Rossier, C.F. Hirjibehedin, Control of single-spin magnetic anisotropy by exchange coupling.
Nature Nanotechnology 9, 64 (2013)

Christopher H. Marrows is Professor of Condensed Matter


Physics at the University of Leeds. He received his PhD there in
1997 and was a Research Fellow of the Royal Commission for the
Exhibition of 1851 from 1998 to 2000. He was the 2011Wohlfarth
Lecturer. His research interests span the fields of nanomagnetism
and spintronics.
Part III
Methods
Magnetic Fields and Measurements
22
Oliver Portugall, Steffen Krämer, and Yurii Skourski

Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1084
Magnetic Field Generation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1084
Permanent Magnets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1085
Electromagnets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1087
High-Field Magnet Facilities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1091
Magnetic Measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1104
Magnetic Field Effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1105
General Technical Principles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1107
Magnetic Field Measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1109
Bulk Magnetic Measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1130
Stray Field Mapping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1139
Calibration and Metrology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1141
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1143

O. Portugall ()
LNCMI-CNRS (UPR3228), EMFL, Univ. Grenoble Alpes, INSA Toulouse, Univ. Toulouse 3,
Toulouse, France
e-mail: oliver.portugall@lncmi.cnrs.fr
S. Krämer
LNCMI-CNRS (UPR3228), EMFL, Univ. Grenoble Alpes, INSA Toulouse, Univ. Toulouse 3,
Grenoble, France
e-mail: steffen.kramer@lncmi.cnrs.fr
Y. Skourski
Hochfeld-Magnetlabor Dresden (EMFL-HLD), Helmholtz-Zentrum Dresden-Rossendorf,
Dresden, Germany
e-mail: skourski@hzdr.de

© Springer Nature Switzerland AG 2021 1083


J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_24
1084 O. Portugall et al.

Abstract

Magnetic fields play a crucial role in manipulating and characterizing the


electronic and magnetic properties of matter. In the present chapter, we discuss
the generation of magnetic fields in a laboratory environment, their measurement,
and the measurement of magnetic properties. As part of the magnet section, we
explain fundamental limitations for generating high magnetic fields, the principal
technical strategies to cope with them, and the implementation of different
concepts in state-of-the-art high-field facilities. The measurement section starts
with a brief review of physical phenomena that can be used to measure a
magnetic field, followed by general technical considerations regarding noise,
signal treatment, and basic sensor requirements. A detailed presentation of
individual sensor types, their operating principle, performance, and application is
given afterward. The subsequent discussion of techniques to measure magnetic
properties essentially refers to the same sensors integrated in dedicated setups.
Stray field methods and questions concerning metrology and calibration are
briefly mentioned at the end of the chapter.

Introduction

Magnetic fields and, more generally, magnetism are associated with a vast number
of physical phenomena as well as a number of misbeliefs and superstitions that
continue to defy scientific evidence right into the twenty-first century. It is therefore
appropriate to recall that magnetism basically just describes interactions involving
two specific properties of matter, charge and spin. Moving charges and spins are
both at the origin of, and affected by, magnetic fields. Their physics thus determines
the construction of magnets as well as magnetic measurements that are the principal
subjects of the present chapter.

Magnetic Field Generation

The co-existence of two distinct properties of matter associated with magnetic fields,
charge and spin, finds its counterpart in the existence of two distinct classes of
magnets, electric and permanent. Each class consists of numerous types and has
its specific practical strengths and weaknesses as far as energy consumption, field
strength, and the modulation of the magnetic field in time and space are concerned.
The specific restrictions of different magnets nevertheless have a common origin,
namely, the limited capability to procure, sustain, and contain the magnetic energy
stored in the field. In this respect, the foremost difference between permanent and
electromagnets is that the first provide the magnetic energy by themselves, whereas
the second depend on an external source.
22 Magnetic Fields and Measurements 1085

Permanent Magnets

Permanent magnets make use of hard magnetic materials that generate a stable
magnetic flux distribution and, consequently, magnetic energy in the surrounding
free space without the need of an external energy source. In terms of applications,
they are widely used as sources of uniform and non-uniform magnetic fields that
can be static or time dependent. With the currently available materials, permanent
magnets are fully competitive with electromagnets for magnetic fields up to 2 T with
record values near 5 T [27]. They are most commonly used as part of a magnetic
circuit composed of permanent magnets with optional soft ferromagnetic materials
that channel the flux to a region of free space where the usable magnetic flux is
generated. The latter is often called the air gap. Therefore, the natural description of
permanent magnets uses the magnetic field strength or flux density B, since flux is
conserved and interactions of electric charges and magnetic moments all depend on
B [27, 39].

Physical and Material Properties of Permanent Magnets


When a permanent magnet is operated in a magnetic circuit in a quasi-equilibrium
state at a given temperature and in the presence of an optional additional magnetic
field, its performance is limited by the fact that the sum of all relevant solid
state interactions plus the energy stored in the magnetic field must represent an
energetically favorable state. Important key parameters of permanent magnets are
their spontaneous polarization Js , their coercive field Hc , and their maximum energy
product (BH )max over the volume of the magnet, which is proportional to the energy
of the magnetic field generated by the magnet in the surrounding free space. This
can be derived from basic magnetostatic relations starting from a situation where the
magnet is part of a magnetic circuit with a magnetic flux density B and a demag-
netizing field H . It can be shown that the energy product never exceeds μ0 Ms2 /4,
where Ms is the saturation magnetization of the material used for the permanent
magnet [39]. Permanent magnets are always based on anisotropic ferromagnetic and
ferrimagnetic materials, since anisotropy is essential to stabilize the magnetization
direction. The basic characterization of magnets involves recording of the M(H )
and M(B) hysteresis loops that provide various macroscopic indicators such as
remanent magnetization, remanent induction, coercive fields for magnetization and
induction, squareness coefficient of the loop, and the maximum magnetic energy
product [39]. These properties originate from the geometry of the magnet (defining
the demagnetization field), the processing (defining the microstructure), as well as
the fundamental microscopic properties of the material like their band structures, the
exchange interactions, and the spin-orbit coupling (the most important contribution
to anisotropy) [156]. Other important parameters are the Curie temperature of
the material and the temperature dependence of the hysteresis loops. Depending
on the type of application, the magnet geometry and the working point (for
static applications) or working range (for dynamic applications) of the magnet in
1086 O. Portugall et al.

the B-H hysteresis loop are chosen. Permanent magnets are either available as
oriented sintered magnets, which show the best performance but are expensive and
complicated to manufacture, or bonded magnets that involve the usage of a non-
magnetic matrix. The latter are cheaper to produce, but their performance is limited.
Materials for permanent magnets can be divided into four groups according to their
maximum energy product and their cost [39]:
Ferrites: Discovered in the 1950s, ferrites are used for traditional and mass
market applications (90% in weight), since they are easily available at low
cost (about 1 US$ per Joule of magnetic energy). They provide good long-
term stability and corrosion resistance. Their disadvantage is their low magnetic
energy product due to their low remanent magnetization.
Neodymium-iron-boron (NdFeB): Discovered in the 1980s, this family is used
for high-performance and miniaturized applications, as they provide the highest
magnetic energy product (above 400 kJ/m3 ; for comparison, lodestone, or natural
magnetite, has a magnetic energy product of 0.7 kJ/m3 ) at still reasonable price
(less than 10 US$ per Joule). Limitations are low Curie temperature and easy
oxidation in air that cause grains to become noncoercive. As these magnets have
the greatest potential for wider applications, intense research is conducted to
improve their properties [156].
Aluminum-nickel-cobalt alloys (AlNiCo): This type of magnets became available
in the late 1930s. Their production cost is almost ten times higher in comparison
to ferrites, yet they provide excellent thermal stability and high Curie tempera-
tures making them ideal for high-temperature and high-precision applications.
Samarium-cobalt magnets (SmCo): SmCo-based magnets were discovered in the
1960s and are high-performance magnets due to their high-temperature behavior
and their compatibility with corrosive environments. SmCo magnets are widely
used in spatial and military applications. Their disadvantage is their high price.

Applications of Permanent Magnets As Flux Sources


Permanent magnets nowadays serve a wide range of purposes including static field
sources for magnetic sensors and measurement techniques requiring a magnetic field
(magnetization, Hall effect, magnetic resonance techniques); traditional industrial
products such as motors, generators, microphones, and loudspeakers; and miniatur-
ized devices in hard disks and sensors for automotive applications. Depending on
their usage and technical principle, three types of permanent magnet applications
can be distinguished:
Static magnetic field sources: Homogeneous and inhomogeneous field profiles
up to 2 T can be obtained by suitable spatial arrangements of permanent
magnets. The magnetic field can be calculated by superposing the contributions
of individual magnets whose mutual perturbation is small due to the rigidity
of their magnetization. In comparison to solenoids, permanent magnets are
better adapted for small systems (mm sized), since the generation of equivalent
magnetic inductions would require impossible electric current strengths or
number of windings [27]. Important applications include Halbach magnet arrays
22 Magnetic Fields and Measurements 1087

that provide highly homogeneous magnetic fields for nuclear magnetic resonance
(NMR) spectroscopy and imaging [133, 31]; single-sided magnets for ex situ
high-resolution NMR spectroscopy [120] and other sensor applications; and
motion and position sensing using magnetic induction or Hall probes [39]. As
permanent magnets generate strong magnetic field gradients, they can also be
used for stray field magnetic resonance imaging in solids (STRAFI). Here, the
magnetic field gradient used for spatial encoding is also generated by the main
magnet, in contrast to conventional magnetic resonance imaging techniques. In
complement to STRAFI methods in solenoids [146,82], their use with permanent
magnets has recently opened the possibility of compact and mobile NMR [31],
as permanent magnets require no external energy source and provide strong
gradients (up to 100 T/m) [27].
Magneto-mechanical applications: Magneto-mechanical assemblies are charac-
terized by forces and torques produced by direct interaction between different
magnets or between magnets on the one hand and magnetic materials on the
other hand. As the magnetic force is given by the spatial gradient of the
magnetic energy, its value depends on the square of the spontaneous polarization
Js . Important magneto-mechanical applications are magnetic gears, bearings,
shock absorbers, magnetic levitation, as well as magnetic separation systems for
paramagnetic and ferromagnetic particles [27].
Electro-mechanical applications: Electro-mechanical applications are charac-
terized by forces and torques produced by interaction between magnets and
devices carrying external or induced electrical currents. The interaction strength
of the forces and torques is proportional to I J , where I is the current and
J is the polarization of the magnet. Electro-mechanical applications represent
the biggest application area, since they comprise motors, loudspeakers, and
actuators involving external currents as well as devices involving eddy currents
as separators for non-magnetic metallic particles [27] and counters for speed or
current consumption [39]. Another important electro-mechanical application is
the use of permanent magnets in beamlines and accelerators in order to guide the
charged particles or to generate curved paths in wigglers and undulators in order
to generate narrowband intense synchrotron radiation [20].

For further discussions of permanent magnets and their applications, see also
 Chap. 28, “Permanent Magnet Materials and Applications,” of this book.

Electromagnets

Electromagnets are current driven and thus governed by the Biot-Savart law relating
the current density j (r) and the magnetic field strength B(r).


μ0 r − r 3 
B(r) = j (r  ) × d r. (1)
4π V |r − r  |3
1088 O. Portugall et al.

As the field varies linearly with the current density, electromagnets are particularly
useful when a magnetic field has to be switched, swept, or tuned. They can also
generate higher fields than permanent magnets and tend to be more practical to
cover larger volumes.
The performance of electromagnets is fundamentally limited by dissipation and
magnetic pressure. These problems as well as advanced technical strategies to
overcome them will be discussed in the next section on high-field magnet facilities.
Here, we limit our attention to common laboratory magnets whose operating
regime gives rise to thermal and mechanical constraints that can be controlled with
relatively basic means. This is the case for resistive magnets providing fields up to
about 5 T and superconducting magnets reaching 23–25 T.
Laboratory magnets are used in physics, chemistry, biology, and material science.
They permit measurements of field-dependent physical properties of matter like
magnetization, Hall effect, magnetoresistance, optical properties, specific heat, ther-
mal transport, and magnetic resonance. Other possible uses include the alignment
or separation of magnetic objects or domains and the manipulation of chemical
reactions and biological processes. Typical industrial applications comprise motors,
loudspeakers, generators, and actuators.
The design of laboratory magnets is subject to application-related requirements
such as the desired maximum field value, homogeneity, a particular spatial field
profile, or geometric constraints like the need of openings for optical access. The
basic approach is to start from the Biot-Savart law (Eq. 1) and to place current
sections such that the desired field profile is obtained. The usage of different circuits
operating at different current levels, of ferromagnetic materials, and of permanent
magnets provides additional degrees of freedom for field shaping. Whereas simple
magnets or magnet assemblies can be calculated by analytical methods [163, 56],
numerical approaches including finite-element techniques have to be used for more
complex arrangements. In addition, constraints given by the magnetic pressure and,
in the case of resistive electromagnets, the energy dissipation may be modelled
numerically.
The most commonly used laboratory electromagnets are resistive and supercon-
ducting solenoids for variable fields, Helmholtz coils providing high homogeneity,
and solenoids equipped with a magnetic core to amplify and bundle magnetic flux.
In the following, these three types will be briefly discussed:

Resistive and Superconducting Solenoids


In solenoids, the current that generates the magnetic field flows along a helical
path on a cylinder surface. In a wider sense, the expression is also used for wire-
wound coils that are composed of successive cylindrical layers, generally of equal
length. Solenoids are called “long” when this length is substantially larger than their
diameter and the magnetic field in the center of the inner region, called the bore,
becomes very homogeneous. They are ideally suited for high-voltage and high-
frequency applications.
Resistive solenoids are mostly used in combination with magnetic materials.
Technical applications include relays as well as linear and rotary actuators, while
22 Magnetic Fields and Measurements 1089

the most typical laboratory devices consist of a pair of electromagnets with


iron pole caps and an air gap. These will be discussed separately below (see
Sect. “Electromagnets with Magnetic Core”).
With the availability of superconducting NbTi or Nb3 Sn alloys, superconducting
magnets have become a standard laboratory tool providing field strengths up to
23.5 T for commercial high-resolution NMR applications (1 GHz) [123] and still
higher fields up to 24 T in dedicated research laboratories [66]. Various research
and development projects are underway to extend this limit by using high-Tc
superconductors [181, 191]. As of 2019, a commercial 28.2 T (1200 MHz) high-
resolution NMR magnet is close to regular operation [152]. Moreover, the high
magnetic field facilities in Tallahassee and Grenoble have successfully operated
prototypes of full-scale high-Tc magnet inserts in superconducting or resistive
background fields providing total magnetic fields of 32 T and 32.5 T, respectively
[111, 160, 161]. In Tallahassee, the potential of high-Tc superconductors has been
furthermore demonstrated by ramping up a scaled-down test insert until it quenched
at 45.5 T [62].
In contrast to electromagnets made of resistive materials, superconducting
magnets exhibit no strong heat dissipation when operated at high magnetic fields
and hence require no high electric power. However, magnetic pressure and the
resulting forces still play a role. In addition, for superconducting magnets, there
are other physical phenomena that limit their maximum field and their operating
conditions. The principal operation constraint is given by the upper critical field
Bc2 (all technical superconductors are type-II), the critical temperature Tc , and the
critical current density jc as shown in Fig. 1. For typical operating temperatures at
4.2 K or near 2.0 K, jc currently limits the maximum available field strengths of
NbTi magnets to 11–12 T and of Nb3 Sn magnets to 23.5 T (cf., Fig. 1). Moreover,
since all technical superconductors are type-II where resistive losses can occur, care
has to be taken that a stable superconducting operating state is maintained under
field changes as well as current and temperature variations [187]. Equally, the
presence of stress induced by magnetic pressure has to be taken into account, as
it can alter the critical current of the conductor. Finally, it has to be ensured that
in case of a breakdown of the superconducting state (quench), the stored magnetic
energy between 0.5 MJ for a NbTi-based magnet operating at 12 T and 4.2 K and
26 MJ for an NMR magnet operating at 23.5 T and 2 K [153] is safely evacuated.
This is achieved by embedding the superconducting wire in a highly conductive
Cu matrix and by appropriate protection circuits including diodes and heating
elements.
It should be noted here that superconducting magnets allow a commercial usage
of high magnetic fields for NMR spectroscopy and magnetic resonance imaging
(MRI). The design of magnets for NMR and MRI is subject to additional constraints
with respect to homogeneity (ΔB/B ≤ 10−9 for a sphere of about 30 mm diameter
for NMR and ΔB/B ≤ 10−5 for a cylinder of about 50 cm diameter and 30 cm
length for MRI) and a time stability of better than 10−8 /h. This is obtained by a set
of superconducting and resistive correction coils (shim coils) and magnet operation
in persistent mode at current densities far below jc .
1090 O. Portugall et al.

Fig. 1 Left: graph showing the relation between upper critical field Bc2 , the critical temperature
Tc , and the critical current density jc of a technical type-II superconductor. The superconducting
phase is limited to points below the surface. Right: comparison of critical current densities (at 4.2 K
unless otherwise stated) available in strands and tapes of more than 100 m. (Adapted from [93])

Helmholtz Coils
Helmholtz coils are a special form of electromagnets that generate a highly
homogeneous field. They consist of a pair of identical solenoids that are thin and
short and placed on a common axis in a distance that equals their radius. These
conditions ensure that the series expansion of the magnetic field in the geometric
center only starts with a fourth-order term.
Main laboratory applications of Helmholtz coils using DC currents are low-field
NMR and electron paramagnetic resonance (EPR) experiments as well as sensor
calibration. Alternatively, Helmholtz coils may also be driven with AC currents in
order to study the effect of time-varying magnetic fields, for example, on biological
systems. Arrangements of three orthogonal Helmholtz coils are furthermore used to
calibrate three-axis sensors and to cancel unwanted magnetic fields and in particular
the earth magnetic field for applications such as zero-field NMR.

Electromagnets with Magnetic Core


Magnetic fields generated by electric currents can be amplified and bundled
with magnetic materials exhibiting a high permeability. This is achieved with
high-susceptibility metals and alloys involving 3d elements like iron. For many
applications, it is desirable to use low-hysteresis materials in order to avoid energy
losses when the field is removed. Annealed soft iron has interesting properties in
22 Magnetic Fields and Measurements 1091

this respect, as it exhibits a high saturation field of 2.16 T at room temperature and
does not remain magnetized at zero field.
Magnetic cores are used in various devices such as laboratory magnets, trans-
formers, electro-motors, magnetic switches, and recording heads. For most of these
applications, the maximum field is about 2 T. However, rare-earth materials have
also been used to boost high-field resistive magnets [144].
A widely used scheme for magnets with magnetic core consists of two solenoid
coils filled with iron cylinders that are aligned on a common axis. Close to the
symmetry plane, the iron cylinders are shaped into pole pieces in order to guide
magnetic flux into the air gap where the experiment is placed. The shape of the
pole pieces can be optimized to provide a suitable compromise between the highest
possible and the most homogeneous magnetic field in the air gap [48, 42, 13, 14].
Magnets of this type are commercially available up to 5 T and require electric power
of the order of 10 kW and water cooling. They are extensively used for a variety
of laboratory applications including most notably electron paramagnetic resonance
(EPR) measurements in the L-, S-, and X-Bands (1–2 GHz/0.06 T, 2–4 GHz/0.12 T,
and 8–10 GHz/0.3 T). For this application, high-class and high-stability power
supplies are required, typically with a stability of less than 10 ppm.
A very useful method for the design of electromagnets with magnetic core is
provided by  the concept of the magnetic circuit analogy to the electric circuit [78].
The term C H · dl = S J · dS (Ampere’s law) defines a magnetomotive force
F (MMF) that takes the role of a “voltage.” In the case of a coil with N turns
and a current I , the MMF becomes NI. The “current” of the magnetic circuit is
represented by the magnetic flux Φ = S B · dS. The equivalent of the resistance
is the reluctance R = F /Φ. For a magnetic circuit in a magnetic material of
permeability μ that has a length l and a cross section A, the reluctance R equals
l/(Aμ). Using this concept, one can find analogs of Kirchhoff’s law in order to
describe circuits containing electromagnets and magnetic materials.

High-Field Magnet Facilities

To obtain the highest possible fields in a practically useful volume requires


substantial amounts of energy, respectively power, as well as the possibility to
control them. In this section, we discuss these limitations together with the principal
technical strategies to overcome them and their implementation in large high-field
magnet facilities.

Performance Limitations and Magnet Classification


The performance of electromagnets is fundamentally limited by dissipation and
magnetic pressure. The first is part of Poynting’s theorem and hence related to
energy conservation in a system composed of electromagnetic fields and particles.
The second refers in the same way to momentum conservation (cf., Maxwell stress
tensor). Both turn out to be proportional to the square of the magnetic field which
explains the substantial technological difficulties in generating high magnetic fields.
1092 O. Portugall et al.

The expression B 2/2μ0 quantifies both the local magnetic energy density
associated with a field B and the pressure exerted on a boundary separating a parallel
field on one side from a field-free region on the other. It is implicitly understood
that such boundary has to be magnetic or contain a net current in order to satisfy the
macroscopic Maxwell’s equations.
Magnetic pressure is primarily a practical quantity that provides a simple upper
limit for the net force a magnet has to sustain. More detailed analyses of local
mechanical constraints make use of the force density that can be evaluated either
via the divergence of the Maxwell stress tensor f = ∇σ where σij = (2Bi Bj −
B 2 δij )/2μ0 or, more commonly, via the Lorentz-force density f = j × B that
refers explicitly to a conductor segment carrying a current density j . In a solenoid,
f gives rise to axial compression and tangential hoop stress j rB where r is the
radius of the current loop. The hoop stress attains its maximum in midplane and
represents the primary limitation for the mechanical integrity of solenoidal magnets.
Under realistic conditions, its magnitude is comparable with the magnetic pressure
of the center field which explains why the latter is often used as a simplified
criterion in design studies. Figure 2 gives an impression of the order of magnitude
of the magnetic pressure and energy density encountered in advanced high-field
magnets.
Dissipation makes it necessary to apply an electric field E that maintains the
current density j in a magnet. The respective power density j · E = j 2 of a
conductor with resistivity ρ gives rise to Joule heating which is quadratic in j
and hence B. Figure 3 gives an idea of the actual power conversion inside a high-
field electromagnet, i.e., the electrical power that has to be provided to maintain
the field and the thermal power that has to be either evacuated (DC magnets) or
temporarily stored (pulsed magnets). The total magnetic energy stored in the coil,

Fig. 2 Conversion scale between magnetic field (lower), pressure, and energy density (both
upper). Arrows indicate the maximum field for different types of magnets (lower) and, for compar-
ison with the magnetic pressure, the ultimate tensile strength of high-strength materials (upper).
Horizontal bars indicate the maximum energy density obtained with electrostatic (capacitors),
electrochemical (batteries), and chemical (fuel) storage. For electrochemical storage, the absolute
maximum obtained in non-rechargeable Li batteries is distinguished from typical values for large-
scale storage in Pb-acid batteries. Although the magnetic pressure and the maximum hoop stress
inside a magnet are not a priori equal, both tend to be similar for most practical magnet geometries
22 Magnetic Fields and Measurements 1093

Fig. 3 Dissipation and the energetic requirements of electromagnets featuring constant current
density, a quadratic cross section parallel to the axis, and a bore size of either 2 cm (solid lines)
or 1/10 (dashed lines) of the outer diameter: (a) required power as a function of magnetic field
and coil size (assuming the resistivity of Cu at 300 K); (b) respective power density and associated
heating if the dumped energy is not evacuated (assuming the heat capacity of Cu at 300 K); (c)
total field/inductive energy stored in the magnet. (a) and (c), respectively, permit the dimensioning
of power supplies for DC magnets and energy storage for pulsed magnets. (b) permits estimates
of the cooling requirements for DC magnets and the maximum field duration for pulsed magnets
depending on the admissible temperature. As a guide to the eye, blue and red circles mark the
approximate parameter space for state-of-the-art DC and pulsed high-field magnets. For real
magnets, power requirements have to be corrected to account for finite filling factors (<100%
Cu content) or the use of high-strength conductors with slightly larger resistivity

B 2/2μ0 integrated over all space, which is just the inductive energy 1/2 LI 2 , is also
shown. The figure permits simple estimates of the operational parameters of DC and
pulsed magnets as shall be discussed below.
In practice, the thermal and mechanical properties of advanced high-field mag-
nets cannot be optimized independently. This statement holds for both the choice of
materials and the general layout. The ideal material for high-field electromagnets
should thus combine high electric conductivity with high mechanical strength,
two properties that tend to be mutually exclusive [162]. The mechanical strength
of a conductor such as Cu can be improved either by work hardening, i.e.,
repeated deformation beyond the elastic limit, or by adding small amounts of
other metals such as Al (Glidcop™) or Ag. Either process, however, reduces the
electric conductivity and hence increases dissipation. The same is true if part of
the current-carrying elements of a magnet are replaced by reinforcing materials,
whence the reduced effective conductor cross section has to be compensated by
higher current densities. Inversely, integrating cooling channels to evacuate the
dissipated heat invariably affects the mechanical properties of a magnet. The design
of high-field electromagnets therefore represents a complex optimization problem
in typically three steps: (a) an initial design choice depending on the basic operating
conditions, the available technical infrastructure, and the envisaged application, (b)
a preliminary concept based on scaling rules and analytic methods [67, 91] that
assume idealized conditions and treat thermal and mechanical constraints separately,
1094 O. Portugall et al.

and (c) extensive finite-element calculations covering the entire magnet assembly
and taking into account the interplay of mechanical, thermal, and in many cases
additional constraints such as the electric behavior of insulating parts or the aging
of all involved materials [180, 5, 171].
Apart from their optimization to obtain the highest field in the largest possible
useful volume, high-field magnets are often subject to additional performance
criteria regarding homogeneity and stability. Relatively good homogeneities are
generally a by-product of the mechanical optimization of magnets as smaller spatial
variations of the field are also necessary to limit stresses. The small but inevitable
radial increase and axial decrease of the saddle-shaped field in the center of a
solenoid can be furthermore flattened by increasing the length h which, however,
implies that more magnetic energy and power is needed. If necessary, correction
coils can also be used to obtain the desired field profile which, however, concerns
almost exclusively NMR applications. Stability usually depends on the power
supply rather than the magnet itself. An exception are superconducting magnets
in persistent mode where spurious losses become important.
Technical details notwithstanding, electromagnets can be divided into categories
depending on their maximum field range, duration, and their approach to overcome
problems related to Joule heating and magnetic pressure. Figure 4 gives an overview.
DC magnets require a robust construction to contain the applied pressure, yet
the principal concern is their isothermal operation. In resistive magnets, heat
has to be constantly evacuated, while electrical energy is reinjected at the same
rate. In superconducting magnets, spontaneous dissipation has to be avoided that
occurs when the kinetic energy of Cooper pairs becomes sufficient to break their
bond and the magnet turns resistive. DC magnets are therefore primarily, if not
exclusively, heat-limited.
Pulsed magnets are operated adiabatically. Their heat capacity buffers a limited
amount of dissipated energy as the field duration is adjusted to keep temperature
changes at a subcritical level. The principal challenge for pulsed magnets thus
becomes the quasi-static containment of magnetic pressure which makes them
mostly force-limited.
Megagauss generators limit the momentum transfer to the magnet in virtually
the same way in which conventional pulsed magnets limit the transfer of
heat, namely, by further reducing the pulse duration. Magnetic pressure is thus
contained dynamically, or, in simple terms, a high magnetic field is generated
before the magnet has time to disintegrate (which it ultimately does). Megagauss
generators are therefore speed-limited.

DC Magnets and Facilities


DC or permanently operated magnets are widely used for scientific and industrial
applications. Highest DC magnetic fields for fundamental and applied research in
physics, chemistry, biology, medicine, and material science are available in large-
scale facilities located at Hefei (HMFL, China), Grenoble (EMFL-LNCMI, France),
22 Magnetic Fields and Measurements 1095

Fig. 4 Performance, operation regime, and categories of high-field electromagnets. The classifi-
cation in heat-, force- and speed-limited magnets indicates a general tendency rather than a hard
fact. In practice, the thermal and mechanical properties of DC and pulsed magnets always need to
be optimized simultaneously

Nijmegen (EMFL-HFML, The Netherlands), Sendai (IMR, Japan), Tallahassee


(NHMFL, United States), and Tsukuba (TML, Japan). Large-scale industrial appli-
cations comprise the confinement of fusion plasmas for energy generation [105]
and magnetic levitation [63]. According to this wide application potential, a large
variety of magnets exist.
High-field DC magnets are based on dissipative high-conductance and high-
strength metallic alloys operated near room temperature. As of 2019, they produce
fields in the 35 to 41 T range. Still higher fields are achieved in hybrid magnets
combining an outer superconducting and an inner resistive magnet. Here the 2017
field record lay at 45 T [104]. In the following, we briefly describe these two types
of magnets.
High-field resistive magnets. Despite substantial efforts to push the limits of
magnets with superconductors toward 30 T, a regular application of more than 28 T
is currently limited to water-cooled resistive magnets requiring dedicated facilities
with tens of megawatts of electrical and equivalent cooling power.
The design of resistive DC magnets is dominated by the necessity to integrate
cooling channels while avoiding any supporting structure or reinforcement that
would further reduce the effective conductor cross section (filling factor). The
magnet structure should therefore be self-supporting, a requirement that is difficult
to satisfy with wire-wound magnets. Over the years, two principal designs have
therefore evolved, Bitter-type magnets and polyhelix magnets. Both concepts are
shown in Table 1 together with a list of key parameters for magnets operated in
large-scale facilities.
1096 O. Portugall et al.

Table 1 State of the art magnets in major high field facilities in 2019. A visual impression of
different magnet types is given on the right. For better comparison a coloured bar in the background

D C - m a g n e t s
Bmax Bore Power Type Fac.
(T) (mm) (MW)
31 50 19 Bitter Ta
31 50 18 Bitter Nij
31 50 24 Polyh. Gre
35 50 24 Bitter Hef
33 32 15 Bitter Tsu
35 32 19 Bitter Ta
36 34 24 Polyh. Gre
38 32 21 Bitter Nij Florida-Bitter principle
38 32 28 Bitter Hef
41.5 32 36 Bitter Ta
36.2 40 14 SCH Ta
38 32 15 Hybrid Tsu
40 (45) 32 28 (>28) Hybrid Hef
45 32 30 Hybrid Ta
42 34 24 Hybrid Gre
45 32 20 Hybrid Nij
P u l s e d m a g n e t s
Buse /Btest Bore FWHM Cool Type Fac.
(T) (mm) (ms) (min)
35/35.8 22 1500 120 mono Wu
43.5 30 1000 240 mono To
64 22 10* 30 flat-t Wu Polyhelix principle
60/65 33 300 120 mono Wu
60 18 35 20 mono To
60/62 28 90 75 mono Tls
65 21 80 35 mono Wu
65 20 10 60 mono Dre
65/67 15.5 20.1 37 mono LA
70/71 24 59 240 mono Dre
70/74 13 55 50 mono Tls
75/85.8 15 4 20 mono To
80/82 13 40 240 mono Tls
83/85 16 5 240 dual Dre
85/90.6 12 23 240 dual Wu
90/91 8 13 90 dual Tls
93/95.6 12 4 240 dual Dre Distributed fibre
95/98.8 8.5 11.8 120 triplex Tls reinforcement
95/100.8 10.5 10.5 180 multi LA
M e g a g a u s s g e n e r a t o r s
Buse Bore FWHM Type Fac.
(T) (mm) (µ s)
To,Tls
150/300 10 5-7 STC LA
650/1200 6-10 2 EMFC To Single-turn coil
22 Magnetic Fields and Measurements 1097

Bitter-type magnets consist of radially slotted copper disks that are vertically
stacked and separated by insulation foils enabling an electrical contact only via a
small segment next to the slit. As the position of the slit and the contact changes
by an azimuth angle along the stacking direction, a series connection of all disks is
established that gives rise to a helical current and hence a magnetic field. Cooling is
ensured by holes that allow a vertical flow of water [15,151]. The shape of the holes
is optimized so as to provide the best possible compromise between heat exchange
and flow resistance. Their radial distribution furthermore anticipates local heating
due to the non-homogeneous current density, while their azimuthal arrangement
close to the bore avoids the radial transmission of forces [9, 10].
Polyhelix magnets consist of a set of (up to 14) concentric cylinders with
individual wall thicknesses of about 1 cm. Each cylinder has a helical slit that is
produced by electroerosion and subsequent gluing with an insulating material. The
entire set of cylinders is connected in series and generates a magnetic field along the
axis of the cylinder. Cooling is ensured by narrow hydraulic channels between two
neighbouring cylinders [149, 33]. Polyhelix magnets are more difficult to fabricate
than Bitter magnets but provide more design freedom as far as the spatial control of
the current density is concerned.
Resistive high-field magnets have reached values between 36 and 41.5 T (cf.,
Table 1) giving rise to stresses up to 400 MPa, a value which exceeds the tensile limit
of Cu (cf., Fig. 2) and comes close to that of other currently available low-resistance
materials featuring conductivities of 45 to 58 MS/m. Although the midplane tensile
stress undoubtedly represents the most crucial limitation for such magnets, their
design also has to pay heed to axial compression that may in turn create shear forces.
Further complications are due to the mechanical robustness of insulating materials
that have to sustain at least part of the applied forces.
Taking into account the finite filling factor of real magnets, i.e., the fraction
of the volume actually occupied by conductor material, one obtains slightly larger
power values than those in Fig. 3. Large facilities therefore operate power supplies
providing 15 to 25 MW giving rise to power densities of typically 2 W/mm3 inside a
magnet. This causes heat flux densities up to 5 W/mm2 at their surface, a value about
one order of magnitude higher than in pressurized vessels of steam generators. The


Table 1 (continued) of each table-cell relates the displayed value to the maximum in the
respective class of magnets. Blue and green colours are used to highlight groups of magnets with
similar features. The last column indicates the facility as explained below. DC-magnet table: SCH
stands for series-connected hybrid. Magnets under construction are in italics. Pulsed magnet table:
Buse /Btest distinguishes fields that are available for users from records obtained with the same
magnet. For nested coils (dual, triplex, multi) the pulse duration (FWHM) refers to the shortest
pulse of the innermost coil. The starred value for flat-top magnets (flat-t, see explanation in the
text) indicates the duration of the plateau. Megagauss table: user fields depend on minimum bore
size. Dre: HLD Dresden [73], Gre: LNCMI Grenoble [92], Hef: HMFL Hefei [71], LA: NHMFL
Los Alamos [110], Nij: HFML Nijmegen [70], Ta: NHMFL Tallahassee [110], Tls: LNCMI
Toulouse [92], To: ISSP Tokyo/Kashiwa [80], Tsu: NIMS Tsukuba [173], Wu: WHMFC Wuhan
[188]
1098 O. Portugall et al.

dissipated energy is evacuated by a constant flow of liquid water at 300 K generating


a typical temperature rise by 30 K. Ideally, the respective cooling power is exactly
equivalent to the applied electrical power, thus permitting continuous operation at
the highest field values. In practice, all but one [150] of the high-field facilities use
pre-cooled water reservoirs to enable operation at electric power levels above the
average cooling power for a limited amount of time. A resistive DC-magnet with its
hydraulic tubing is depicted in Fig. 5.
Hybrid magnets. Although it would be technically possible to generate fields of
the order of 50 T in resistive magnets [5], controlling the thermal and mechanical
constraints would require a substantial increase of their size and a reduction of
their filling factor to provide additional space for cooling and mechanical support
[8]. Therefore, higher DC magnetic fields are generated with hybrid magnets that
combine a resistive and a superconducting part. Table 1 provides an overview of
operational and projected hybrids. Further details can be found in the literature
[11, 21, 49, 114, 113, 129, 185].
Hybrid magnets are able to generate magnetic field strengths above 30 T with
only 10 MW of electric power and above 40 T with 20 MW. The resistive and
superconducting parts are usually operated by different power supplies. This is due
to the preferred operation mode for hybrids: the superconducting part is first ramped
to maximum field and kept there; then, field ramps are only performed with the
resistive part. Another reason for separate power supplies is the different electric
power consumptions and characteristics of the two types of magnets: the resistive
part typically consumes 20 MW and exhibits low inductance values (below 50 mH),
whereas the superconducting part has a large inductance (up to 3 H) and requires
only a very small fraction of the total power. However, there also exists a special
variation of the hybrid design, the so-called series-connected hybrid, where the
superconducting and resistive parts are connected in series to the same power supply
[36]. This design provides enhanced stability and resolution, since fluctuations of
the resistive magnet and the power supply are damped by the superconducting part.
Series connected hybrids therefore enable high-resolution NMR in the 30–40 T field
range [55].
Owing to the critical field that limits superconductivity, hybrid magnets always
consist of a superconducting outsert and a resistive insert. Since the field of the
outsert exerts additional forces on the insert, the field strength of the latter is
reduced with respect to its maximum value in stand-alone operation. Mechanical
stability is also a problem for the superconducting part, as strong magnetic forces
between the insert and the outsert may occur in case of asymmetries due to imperfect
mounting or partial defects in the resistive part. The design and operation of hybrid
magnets furthermore requires special considerations regarding the stability and
safety of the superconducting part in the presence of a rapidly ramped resistive
insert.
Current hybrid projects are aiming to obtain magnetic fields up to 45 T using
a combination of low-temperature superconductors and resistive magnets. Future
plans aim to reach a magnetic field of the order of 60 T. However, using present-
day technology, such field values would require large electric power (40–50 MW)
and a superconducting outsert operating near 20 T with a stored energy of the order
22 Magnetic Fields and Measurements 1099

Fig. 5 (a) Historical image of the Bellevue electromagnet built in 1928 to generate 5–7 T (©CNRS
photothèque, with kind permission); (b) modern triplex-pulsed magnet with 700 mm outer diameter
producing close to 100 T (©LNCMI, with kind permission); (c) hydraulics tubes around the chassis
of a resistive DC magnet (©Steffen Krämer); (d) 50 MJ capacitor bank for generating pulsed fields
(©HZDR, with kind permission)

of 1 GJ [129]. Consequently, next-generation hybrid magnets will have to involve


high-temperature superconducting materials with their capacity to operate beyond
30 T [111, 160, 161, 62]. Design studies are underway to reach this ambitious goal
[11, 128].

Pulsed Magnets and Facilities


Pulsed magnets [67] provide higher fields than DC magnets, are cost-efficient in
terms of investment and operation, and can be constructed with relatively basic
technical know-how. In Western Europe alone, small pulsed magnet facilities have
been, or are, operated in at least 15 different laboratories. The creation of large
infrastructures is a recent phenomenon. As of 2017, fully operational facilities
making routine use of energies in the 10 MJ range exist in Dresden (EMFL-
HLD, Germany), Los Alamos (NHMFL, United States), Toulouse (EMFL-LNCMI,
France), and Wuhan (WHMFC, China).
As outlined before, in pulsed magnets, the dissipation problem is solved by
reducing the field duration. For all practical purposes, this provides satisfactory
solutions despite the fact that shorter pulse durations also give rise to additional
heating by eddy currents or, in an alternative description, the finite current-
penetration depth that effectively reduces the available conductor cross section. This
1100 O. Portugall et al.

penetration depth is related to, but not identical with, the ordinary skin effect that
only applies for stationary oscillations.
More recently, pulsed magnets have also been equipped with cooling channels,
not to evacuate heat during shots but in between [53]. This gives rise to considerably
enhanced duty cycles which represent an important factor for applications in user
facilities.
Realistic estimates for the heating of pulsed magnets can be obtained by integrat-
ing both sides of the expression j 2 (t) dt = D(T ) cp (T )/ρ(T ) dT that associates
the application of a current density j for a short interval dt with a temperature
rise dT in a conductor that is characterized by its temperature-dependent mass
density D, specific heat cp , and resistivity ρ. For most conductors, this so-called
action integral becomes more favorable at low temperatures which is one of the
reasons why advanced pulsed magnets are pre-cooled with liquid nitrogen. A base
temperature of 77 K also leaves a larger margin with respect to the relatively limited
heat tolerance of many insulating materials. The heating of pulsed magnets during
a shot can easily exceed 100 K.
In complete analogy to resistive DC magnets whose heat limitation motivates
the integration of cooling channels, pulsed magnets require reinforcing elements
to contain the applied magnetic pressure. First-generation pulsed magnets were
therefore reinforced externally. Their design was based on analytic calculations and
the destructive testing of downsized prototype coils followed by the application
of scaling rules to build larger specimens to be operated below the destructive
limit. The main shortcoming of this approach is that the magnet is regarded as
a mechanical continuum rather than a complex structure composed of helical
windings, isolations, reinforcements, and impregnations which cannot be easily
scaled simultaneously.
With one exception [89], modern pulsed magnets or pulsed magnet inserts
attaining the highest available fields are all based on the same basic concept of
distributed fiber reinforcement [17]. The principle is that of a wire-wound magnet
whose conducting layers are individually stabilized with layers of high-strength
fibers, a technique that has the additional benefit of electrically insulating adjacent
conductor sections. The material of choice is a synthetic polymer fiber commonly
known as Zylon™ whose tensile strength comes close to that of carbon fiber albeit
without the disadvantage of being conductive. The thickness of each reinforcement
section is adjusted such that neither a substantial transmission of forces between
successive layers occurs nor a separation when the coil is elastically deformed.
In recent years, the perfection of finite-element methods has shifted the design
focus from the comparably simple question of how to contain maximum stresses
in midplane to problems related to aging and local forces caused by the imperfect
cylinder symmetry of real magnets. In the latter respect, current terminals and the
transition points at either end of the magnet where wire is guided from one layer to
the next are potentially weak spots. Aging, on the other hand, can be due to either
repeated deformation beyond the elastic limit if a magnet is operated close to its
maximum, or thermal cycles and, in some cases, also the electric or mechanical
deterioration of insulating materials.
22 Magnetic Fields and Measurements 1101

Table 1 gives an overview of key parameters of advanced pulsed magnets.


The relatively large variation of parameters such as bore size, pulse duration, and
cooling time reflects technical limitations of available energy sources as well as
deliberate choices to account for specific experimental requirements. Nested coils
marked “dual,” “triplex,” or “multi” resemble DC-hybrid magnets insofar as their
architecture permits different technical choices closer to the bore where fields are
highest and further away where most of the energy is stored and needed. In nested
systems, the innermost coils thus dispose of the strongest reinforcement which
reduces their filling factor, i.e., the relative amount of conductor material, and hence
their heat capacity. As a consequence, they require shorter pulse durations and are
powered independently. An example of a pulsed triplex-coil with 3 independent
coaxial current terminals is shown in Fig. 5.
With few exceptions, pulsed field installations use capacitor banks as energy
source, cf., Fig. 5. Figure 6 shows the operation principle depending on whether a
simple serial RLC arrangement with a primitive closing switch is used or a modern
setup with a thyristor switch and diode crowbar. The use of solid-state devices makes
the switching practically noise-free and allows for some control of the discharge.
A full crowbar circuit thus avoids a voltage reversal that ultimately deteriorates
capacitors and thyristors, reduces the heating of the magnet, and produces a smooth
field decay that is preferable for field-dependent measurement applications.
Losses notwithstanding, the energy stored in the capacitor bank 1/2 CU 2 provides
the magnetic energy, cf., Fig. 3c, and hence limits either the coil size for a given peak
field or vice versa. The capacitance C imposes further restrictions since together
with the inductance of the electrical circuit – primarily the coil – it determines the
pulse duration and hence the heating as demonstrated in Fig. 3b. The application of
high voltages is therefore unavoidable. In practice, capacitor banks for conventional

Fig. 6 Voltage, current, and coil heating for different configurations of capacitor-driven pulsed
field generators. Violet, generator with a simple closing switch S and no crowbar CB; Red, thyristor
switch but no crowbar; Orange, modern crowbar circuit. The thyristor-and-crowbar option has the
advantage that the capacitor bank is discharged at the end of the pulse. The smooth decay is also
favorable for measurement applications, and the heating can be tuned via the crowbar resistance
1102 O. Portugall et al.

pulsed magnets are limited to 25 kV, a range where corona and other partial
discharge phenomena can still be controlled with relatively simple technical means
and where self-healing metallized film capacitors for pulsed power applications are
widely available on the market.
Although state-of-the-art capacitor banks in large facilities feature energies
exceeding 10 MJ, a variety of smaller sizes exists ranging from transportable
systems in shipping containers over mid-sized installations for special applications
[2, 52, 81] down to tabletop and even microscopic generators [96]. Special architec-
tures have also been developed to produce repetitive [79] and flat-top magnetic fields
[85]. As far as alternative energy sources are concerned, inductive storage has been
successfully tested in the past [6], but the project has been discontinued. Flywheel
generators originally built for other high-power applications have been recycled by
some facilities in order to obtain energies approaching the gigajoule range and the
possibility to implement controlled waveforms [147, 90].
Although generating the highest possible fields for scientific research remains at
the center of development efforts, pulsed magnet technology has started to expand
into other directions. Recent innovations for scientific purposes include split coils
providing lateral access in order to combine high fields with intense radiation
sources [40, 81] and pulsed dipole magnets for investigating the quantum vacuum
[7]. Pulsed magnets have also found their way into industrial applications where
they are used for magnetizing permanent magnets, electromagnetic forming, and
magnetic pulse welding.

Megagauss Magnetic Fields


Megagauss generators provide magnetic fields well beyond the current techno-
logical limit for conventional pulsed magnets, i.e., 100 T or 1 megagauss in cgs
units. Both their production and their use in scientific applications require highly
specialized equipment that is available in only a few laboratories worldwide. In
2019, Megagauss fields for scientific research were generated on a regular basis in
Los Alamos (NHMFL, United States), Tokyo (ISSP, Japan), and Toulouse (EMFL-
LNCMI, France).
Often coined destructive techniques because coils are inevitably destroyed,
Megagauss generators make explicit use of inertia that resists the displacement
of conductor elements by a magnetic force. Interestingly, Newton’s historical
distinction between objects that are either initially at rest or in a state of uniform
motion finds its counterpart in two distinct techniques for generating Megagauss
fields.
Single-turn coils have been reported to produce fields up to 355 T [154], yet their
principal interest lies in their capacity to generate 150 to 200 T in a useful diameter
of 10 mm without causing excessive peripheral damage [68,109,125,103]. The latter
applies for thin-walled single-turn coils that basically consist of a rectangular copper
strip bent to form a single coil winding (Table 1). When the coil disintegrates, the
magnetic pressure in the bore propels fragments radially outward, thus protecting
equipment in the center. As of 2015, single-turn coils with almost identical geometry
22 Magnetic Fields and Measurements 1103

and typical dimensions of 2h = 2a1 = 10 ∼ 15 mm and a2 − a1 = 2 ∼ 3 mm were


operated by three facilities worldwide to perform regular scientific experiments.
Although their physics is characterized by a complex interplay of electrody-
namic, thermodynamic, and mechanical processes [67, 91], the basic working
principles of single-turn coils can be derived relatively simply. Most importantly, a
relevant timescale can be determined by assuming the magnetic pressure associated
with a sinusoidal field, neglecting all elastic forces and integrating the equation of
motion dF = p dA = dm r̈ for the radial displacement r of a conductor element.
A sensible limit of <10% for the effective dilation at peak field then identifies the
risetime for a copper coil with 10 mm inner diameter and 2 to 3 mm wall thickness
to be of the order of 1 μs.
The short duration puts heavy constraints on the design of capacitor banks for
single-turn coils as both the capacitance and inductance have to be minimized
simultaneously. Although the small size of the coil requires very modest energies
of the order of 100 to 200 kJ, operating voltages still have to be raised up to 40
to 60 kV. As a consequence, insulating and conducting parts have to be carefully
shaped and finished to avoid partial discharge. Insulation also has to be kept thin to
avoid unnecessary inductance contributions. Furthermore, the intense peak currents
of typically a few mega-amperes require a careful design of electrical contacts to
prevent deterioration by arcing.
Attempts to improve the performance of single-turn coils include the use of thick-
walled specimens, heavier conductor materials, and larger diameters. The latter give
rise to longer pulse durations but also require substantially larger energy sources to
compensate for the increased volume [90]. Heavier conductor materials such as
tungsten have been successfully tested [109] but are impractical for routine use
due to their higher cost. The original use of thick-walled single-turn coils has been
largely abandoned due to extensive damage created inside the bore. The latter is
a consequence of the finite penetration depth that concentrates the current flow
and magnetic force near the edges and the inner surface of the coil [109, 125].
As the latter is pushed against a heavy outer mass, it bounces back into the bore
instead of moving radially outward. In addition, the inner surface of the coil partially
sublimates due to Joule heating. At this point, the moderate dilation experienced
by a thinner coil is beneficial as it permits the evacuation of the generated vapor.
The increasing importance of sublimation as opposed to disintegration is believed
to be responsible for practical limitations of single-turn coils [125]. Contrary to
what might be expected, the breaking of the coil once the field descends has little
effect as the current continues to flow by plasma discharge. Single-turn coils also
feature a better-than-expected homogeneity because of the inhomogeneous current
distribution.
Flux compressors generate higher fields than single-turn coils albeit at the
expense of substantially larger peripheral damage including the inside of the bore.
Their basic principle is that magnetic flux can be auto-trapped in a conducting
loop, called liner, if the latter is compressed with such speed that the flux density
undergoes changes limiting its own penetration into, and diffusion through, the
electromagnetic boundary [67]. Ideally, the magnetic field thus experiences an
1104 O. Portugall et al.

Fig. 7 Electromagnetic flux compression principle: red arrows indicate magnetic flux, blue arrows
the current in the single-turn coil (STC, grey) and the copper liner (orange). (a) Flux generated by
the STC induces eddy currents in the liner and diffuses only partially through to the center. The
liner is compressed by the pressure associated with the larger flux density between its outer surface
and the STC. (b) As the liner starts to implode, flux in its center is compressed, while the flux
outside is decompressed; when the flux densities inside and outside are equal, the sense of the net
current in the liner and the pressure are inverted; however, the liner is now in full motion and thus
carries on through its own inertia. (c) The inertial movement continues to compress flux captured
in the center until the kinetic energy of the liner is fully converted into field energy

amplification that is exactly proportional to the reduction of the area enclosed by


the liner. The corresponding increase in magnetic field energy is compensated by
the loss of kinetic energy of the liner.
Figure 7 visualizes the principle of an electromagnetic flux compressor that
makes use of the magnetic field generated by an oversized single-turn coil, both
to create the so-called seed field inside the liner and to drive the implosion [26].
Alternatively, the seed field can also be generated by an additional magnet, and
the compression can be achieved by applying high explosives which increases the
damage but pushes the field limit from currently 1200 T for electromagnetic devices
[108] to more than 2000 T [19].
In practice, the design of flux compressors represents a complex optimization
problem since the thermodynamic state of the liner changes during the compression
and thereby influences the electromagnetic properties governing the diffusion pro-
cess [67]. Further complications are associated with mechanical inhomogeneities
and inadequate geometries that can provoke liner instabilities and hence an inhomo-
geneous or incomplete compression [99].

Magnetic Measurements

Measuring a magnetic field means to observe its effect on charge and spin. The
surprising diversity of magnetic field sensors results from the transformation of this
initial influence in three steps: Firstly, atomic, molecular, and solid-state interactions
alter the behavior of charges and spins and hence the way in which they respond
to an applied field; secondly, a given microscopic phenomenon affects different
22 Magnetic Fields and Measurements 1105

macroscopic state variables and response functions; and thirdly, few sensors are
simple transducers that convert the applied field directly into a standard voltage
output.
In this section, we start with a brief outline of magnetic field effects that can
be exploited for sensor applications. The principal mechanisms shown in Fig. 8
reappear in diagrams explaining the operation principles of different sensor types
in the subsequent paragraphs.

Magnetic Field Effects

Upmost in the hierarchy of influences affecting the behavior of charges and spins are
atomic and solid-state interactions that regroup both in three categories. Quasi-free
charge carriers that give rise to Amperian currents exhibit single-particle behavior
while being characterized by their linear momentum. Likewise, magnetic moments
refer to charges or spins that behave individually but whose prominent dynamic
feature is their angular momentum. The combined effect of Coulomb repulsion and
Pauli principle finally generates effective exchange interactions that are at the origin
of collective magnetism.
Formally, the kinetic momentum of a system composed of charged particles
and electromagnetic fields, p − eA, represents the unique origin of magnetic field
effects in both classical and quantum theory. Here p, e, and A denote the canonical
momentum, charge, and vector potential. Spin-related field effects occur when
the expression is introduced into Dirac’s equation. In non-relativistic theories, the
respective terms are introduced ad hoc.
The well-known classical magnetic forces and equations of motion can be
deduced from p − eA through the Lagrange-Hamiltonian formalism. As part of
the Schroedinger or Heisenberg equations for quasi-free particles, p − eA produces
additional quantum√ effects, namely, a spatial modulation characterized by the
magnetic length h̄/eB that is at the origin of Landau quantization and Aharonov-
Bohm like quantum interference effects. Under realistic conditions, B < 103 T, this
is the only case where a magnetic field is quantizing in the sense that it generates
new quantum numbers.
For magnetic moments an external field only lifts the orientational degeneracy,
unless it reaches B ∼ 105 T, and the magnetic length is reduced to the order of the
atomic radius. All other cases are adequately described by the Zeeman Hamiltonian
and higher-order corrections that are the result of p − eA being expanded in terms
of angular momentum operators. Although magnetic fields do not generate new
quantum numbers in this case, their competition with, for example, a weak spin-
orbit interaction may nevertheless tip the balance between different sets of angular
momentum eigenstates. The Zeeman shift may also change the ordering of quantum
states and thus affect the ground state of a system.
In exchange-coupled systems, the situation is similar, except that the field
also has to compete with existing directional constraints due to the spontaneous
alignment of magnetic moments. The result varies substantially with the type of
1106 O. Portugall et al.

... Flux &


alignm.
effects Zeeman
energy

Larmor
Gradient precess. Class.
Torque
force cyclotron
(micro)
motion
Hall
effect

Magnetic
moment Lorentz Force on
force conductor

bound

Mobile
Spin Charge free
charge
EMF
Faraday
induct.

exchange coupled

Magnetic
Collective length
magnetism
Landau
quant.

S-ferro: Quantum
H-ferro: domain interf.
object dynam. Flux
dynam. concentr.

Spin-
Magneto- polariz.
Macrosc. striction
motion

Fig. 8 Scheme of effects that may be used in magnetic sensors. The formation of field-sensitive
media is represented inside the shaded area. The first bubbles outside indicate how magnetic fields
affect the respective equations of motion. H-ferro and S-ferro denote hard and soft ferromagnets.
The first react as macroscopic objects to a magnetic field, while the second change their internal
structure. Gray bubbles represent effects that are not generally used for sensing purposes
22 Magnetic Fields and Measurements 1107

magnetism and is in general subject to threshold and other nonlinear effects. In the
following, we limit our attention to hard and soft ferromagnets. In the first case,
the alignment force is effectively transferred to the macroscopic object, while in the
second, it affects the orientation of domains and thus gives rise to magnetization
changes.

General Technical Principles

The choice of a physical principle for a particular sensor application is governed by


criteria that fall into three groups:
Signal properties versus sensing requirements: Of foremost importance are
properties of the measured field, i.e., its strength, frequency bandwidth, and
spatial variation, and the required accuracy and precision of the measurement.
In addition, sensors may need to detect the direction of the applied field (vector)
or just its amplitude (scalar).
Technical environment: Most sensors are subject to further constraints related to
the environment in which they operate. Typical examples are restricted volumes,
limited access to electrical power, or the need to comply with specific chemical,
electromagnetic, or thermodynamic conditions.
Fabrication and cost: Raw materials, production techniques, and maintenance
determine the economic investment for producing and using a sensor and
hence its suitability for mass fabrication (e.g., read-heads, position sensors),
small series production of hi-tech devices (e.g., magnetic anomaly detectors,
magnetoencephalographs), or the design of unique prototypes in science and
research.
Different operating principles and applications notwithstanding, magnetic sen-
sors tend to suffer from the same problems and often apply similar concepts to
overcome them. Prior to the subsequent review of individual sensor types, we
briefly discuss these common features including a short reminder of relevant sensor
parameters.

Technical Parameters: Accuracy, Precision, Sensitivity, Resolution, and


Responsivity
Accuracy and precision are the primary, most general, sensor specifications. Both
are expressed in units of the measured quantity, in our case T. Not to be confused,
precision characterizes the scattering of values when the same quantity is measured
several times, whereas accuracy quantifies the deviation between the observed mean
value and the true signal level.
Accuracy depends primarily on a sensor’s calibration and physical principle
which may or may not be suitable for absolute measurements. This problem will
be addressed in Sect. “Calibration and Metrology”. Precision, on the other hand, is
dominated by technical issues such as a sensor’s intrinsic noise level, its nonlinearity
and hysteresis, and its offset and drift. It is related to sensitivity that quantifies
1108 O. Portugall et al.

the input signal required to produce a significant output compared to artifacts and
perturbations. The ubiquitous effect of noise in magnetic √ sensors, for example, is
characterized by a noise-equivalent field expressed in T/ Hz.
Slightly less specific, resolution is defined as the smallest input variation that
produces changes in a sensor’s output. Originally a key parameter for instruments
relying on discrete internal references (e.g., frequency in NMR) or featuring digital
output, resolution is also used in a wider sense to characterize a sensor’s lower
detection limit.
Sensors operating as transducers in larger setups where a certain output level
is mandatory are frequently characterized by their input-output signal ratio or
responsivity. Unlike all previously discussed parameters, responsivity provides no
real indication of relevant detection limits as it does not take into account intrinsic
perturbations.

Common Problems: Noise


Predictably, sensor noise can be of electric or magnetic nature. Johnson noise
and shot noise are frequency-independent perturbations, the so-called white noise.
They respectively refer to the thermodynamic state of a conductor and the process
of charge transport. Johnson or thermal noise arises from the excitation, relax-
ation, and
√ erratic motion of charge carriers. It gives rise to voltage fluctuations
VJN = 4kB T R f where kB T , R, and f denote the thermal excitation energy,
resistance, and frequency interval. Shot or Poisson noise is a general statistical
phenomenon that occurs when a process√is composed of discrete events. A current
I thus produces fluctuations ISN = 2eI f because it relies on the motion
of individual electrons. In most practical cases, the effect of shot noise is less
pronounced than that of Johnson noise. A notable exception is systems featuring
heterojunctions were shot noise can become dominant. As a rule, white noise is best
suppressed by limiting the measurement bandwidth f .
Burst or random telegraph noise and in particular 1/f noise are the dominant
perturbations at low frequencies. In the simplest case, burst noise originates from
the arbitrary switching between two states in a bistable system. The frequency of
the respective noise is limited by the average time the system remains in a given
state. Pink or 1/f noise is a ubiquitous phenomenon in Nature characterized by its
inverse frequency dependance. In conductors, it is manifest as resistance variations
that have their origin in the dynamic fluctuation of scattering centers, traps, and
other imperfections. Magnetic materials, on the other hand, are subject to 1/f
noise originating from fluctuating domain structures. Both burst and 1/f noise may
be suppressed with the aid of phase-sensitive detection (PSD) techniques which,
however, requires the capacity to modulate the measured signal. Magnetic 1/f
noise may also be eliminated by controlling the domain structure of the respective
materials.
Barkhausen noise is another magnetic perturbation that accompanies magnetiza-
tion changes in soft ferromagnets. It consists of discrete jumps that reflect the sudden
orientation of entire domains or groups of coupled domains. The magnetization
22 Magnetic Fields and Measurements 1109

dynamics are material dependent and ultimately limit the frequency range that can
be covered. This is particularly important in flux concentrators.

Common Solutions: Flux Concentration, Modulation, and Null Detection


Flux concentrators play a crucial role in magnetic sensor technology. Here, a weak
applied field generates additional flux by tuning the demagnetizing field of a soft
ferromagnet. Depending on the direction of the field and the geometry of the flux
concentrator, a suitably positioned sensor thus captures a larger amount of flux.
Flux concentrators are therefore extensively used as passive amplifiers to measure
small fields in limited volumes. The principal factors determining their quality are a
large permeability, small hysteresis, high Curie temperature, small magnetic noise,
negligible magnetostriction and high resistivity to prevent eddy current losses in
time-dependent fields. Their design and adaptation to specific sensors also have
to take into account the effect of orthogonal fields. Inevitable disadvantages are
their saturation and magnetization dynamics that limit the upper field and frequency
ranges and their behavior with respect to orthogonal fields. The stray field of a flux
concentrator may also affect the applied field via its source if the latter is small and
within range.
As mentioned before, modulation techniques are most efficient to cope with
1/f noise. Modulation is relatively simple in sensors that actively probe magnetic
fields via their – direct or indirect – effect on currents or electromagnetic radiation
whose time dependence can be controlled. Modulating the effect of an applied
field on a passive sensor is generally more difficult due to the need of a magnetic
shutter. Fluxgate sensors thus force a soft ferromagnet periodically into saturation
to modulate its effective permeability with respect to an applied field. A more
direct and generally applicable method consists of mounting a flux concentrator
on a mechanical oscillator. This solution has gained importance with the advent
of micro-electro-mechanical systems (MEMS) whose miniature scale stands for
compact sensor architectures, higher modulation frequencies, and reduced energy
consumption. Prototype modulators providing frequencies up to 15 kHz have been
reported [44].
Another problem of many sensors, especially those relying on magnetic mate-
rials, is their intrinsic nonlinearity. A standard solution for this problem is null
detection where an integrated bias coil compensates the applied field or otherwise
shifts the sensor’s working point into its linear operating range.

Magnetic Field Measurements

Magnetic sensors have been reviewed in books [141, 175] and articles [43,
94, 143, 176] often with emphasis on specific techniques, applications, or recent
advances. The present section presents a brief overview over the principal sensor
types, their measurement principles, performance, and practical importance. Further
information may also be found in  Chap. 31, “Magnetic Sensors,” of this book.
1110 O. Portugall et al.

Table 2 Operation and performance limits for different sensor technologies. The parameter ranges
refer to sensor families as opposed to individual sensors whose specifications are generally much
more restrictive
Type Range [T] Bandlimit [Hz] Noise [T/Hz1/2 ] Practical importance,
specifications & references
Induction coil 10−13 –10+3 10−3 –10+9 10−13 @ 20 Hz, Highly adaptable, flexible
air design;
10−12 @ 1 Hz, easy to fabricate; [140, 174]
ferro
Fluxgate 10−11 –10−3 <10+3 10−12 Standard solution for
10−9 –10−12 T;
commercial products;
[127, 139, 142]
Hall effect 10−5 –1050 <10+6 10−7 Standard solution for
>10−3 T;
commercial products,
chip-integration
GMR, TMR 10−9 – <10+9 10−8 Standard solution for
AMR 10−3 <10+7 10−9 –10−3 T;
commercial products,
chip-integration; [167, 101]
SQUID 10−15 –10−2 10+9 @ 4.2 K 10−15 @ 4.2 K Standard solution for
10+4 @ 77 K 10−13 @ 77 K <10−12 T;
commercial products;
[170, 16]
NMR 10−10 –50 <10+2 10−6 , standard Absolute
10−12 , measurement/calibration
optimized standard;
sensitivity gain with
Overhauser effect (×10−1 )
or ESR/optical pumping
(×10−2 ); [76, 59, 61, 112]
Faraday effect 10−8 – <10+9 10−12 Electromagnetic
10+3 compatibility & remote
sensing;
10−11 T in advanced
laboratory tools
Lorentz force 10−9 –1050 <10+3 10−9 Emerging technology,
& other mostly Lorentz-force;
MEMS power-free sensing with
magnetostriction, torque;
[43, 69]
Magnetoelectric 10−10 –10−5 10−9 Emerging technology,
power-free sensing;
different device sizes;
[98, 179]
22 Magnetic Fields and Measurements 1111

Table 2 lists operation and performance limits for different sensor technologies.
The device specifications of individual sensors are generally much more restrictive
as both their size and architecture may differ substantially. This is particularly true
for induction coils. The fact that different scientific domains have produced their
characteristic architectures also explains why certain sensor types are assessed quite
differently in literature.

Induction and Fluxgate Sensors


Induction coils [29, 141, 174], also called search or pick-up (PU) coils, are a direct
implementation of Faraday’s law. Here, the presence of free charges on a contour,
i.e., the coil, creates the boundary condition that permits integration of the Maxwell-
Faraday equation ∇ × E = −∂B/∂t. As charges are displaced along the contour
by the electromotive force (EMF), the coil terminal exhibits a voltage that reflects
changes of the enclosed flux, V = −∂Φ/∂t. The optional use of soft-ferromagnetic
cores as flux concentrators increases the effect as the coil now senses both the
external flux and flux generated by the core magnetization.
The working principle of induction coils is illustrated in Fig. 9. Table 3 lists the
respective transfer function, i.e., the ratio between the measured voltage and the
field strength assuming that the latter is homogeneous across the coil area. If the
necessary time dependence is not an intrinsic property of the applied field, it may
be generated by moving or rotating the coil.
The resolution
√ of induction sensors is limited by thermal noise depending on coil
resistance as √R. Improving an air core coil means to increase its area such that
Anew /Aold > Rnew /Rold . Larger coil diameters are most efficient as can be seen
by doubling a length of wire that forms a single winding: whereas Rnew = 2Rold ,
the radial expansion gives Anew = 4Aold thus improving the resolution by 23/2 .
Adding another winding in the axial direction produces the same Rnew but reduces
the gain to 21/2 as Anew = 2Aold . No gain at all is achieved if coil dimensions are

Soft-ferro core Induction coil

dB 0 Faraday
V
dt induction

dB 0 dB μ Faraday
Magnetization V
dt dt induction

Fig. 9 Working principle of induction coils with air (top) and soft ferromagnetic core (bottom).
B 0 is the external field, B μ the field associated with the core magnetization, and V the induced
voltage
1112 O. Portugall et al.

Table 3 Transfer function of induction and fluxgate sensors. For induction sensors a harmonic
time dependence of either the external field B0 or the coil area A was assumed to illustrate the
frequency dependence. V is the readout voltage, μr the core permeability, N the demagnetization
factor, ω the angular frequency, and i the imaginary unit representing a π/2 phase shift. The
fluxgate equation refers to the residual permeability μr that oscillates between 0 and μr
Type Transfer function Modulation
Air core V /B0 = iωA B0 ∼ eiωt (or A)
Magnetic core V /B0 = iωA μr [1 + N (μr − 1)]−1 B0 ∼ eiωt (or A)
dμr  −2
Fluxgate V /B0 = A dt (1 − N ) 1 + N (μr − 1)

fixed and the wire cross-section is cut in half to accommodate a second winding,
hence Rnew = 4Rold .
Whenever small coil areas are mandatory, the resolution of induction sensors can
likely be improved with a ferromagnetic core producing an approximately constant
gain of ∼ N −1 as long as N μr  1 (cf., Table 3). Signal amplifications of close
to 103 may therefore be obtained while avoiding, at least partially, the nonlinearity,
temperature dependence, and Barkhausen noise associated with the permeability
μr . In return, the saturation and magnetization dynamics of ferromagnetic cores
respectively give rise to a material-dependent reduction of the upper field and
frequency limits.
Air core coils exhibit substantially larger upper bandlimits caused by the onset
of nonlinear behavior close to their resonance frequency. The latter arises from
parasitic capacitances between close-packed windings or the wiring between coil
and voltmeter. The input impedance of preamplifiers or low-pass filters used to
integrate transient signals may alter the frequency response further and thus have
a strong effect on sensor performance [29, 141, 174]. Since the output of induction
sensors depends on frequency, their lower band offset and resolution are correlated.
As a rule of thumb, air core coils provide extreme sensitivities, field, or frequency
ranges, whereas ferromagnetic cores are the best option for optimizing resolution in
a restricted volume.
Their simple architecture makes induction sensors a popular choice in science
and industry. They use basic materials (wire), are easy to manufacture (winding),
and do not require sophisticated readout technology (voltage measurement). Air
core coils can be made extremely large for applications such as traffic control
or geomagnetic measurements. Coils with 2 m diameter, 16,000 turns, a lower
frequency limit of 0.004 Hz and a resolution of 10−12 T have been reported [164].
Pick-up coils for measuring Megagauss fields, on the other hand, consist of a
single winding with less than 3 mm diameter and an upper frequency limit of
0.1–1 GHz that is determined by readout electronics rather than the coil itself.
The design flexibility of induction coils has also given rise to a variety of special
developments such as gradiometers, Rogowski coils for current sensing, or packs
of three orthogonal coils for vector measurements. Gradiometer coils, i.e., pairs of
compensated induction coils, can reach sensitivities of 10−13 T while remaining
compact enough for biomedical applications such as magneto-cardiography [46].
Planar thin-film coils (e.g., [58]) are an interesting alternative for producing simple
22 Magnetic Fields and Measurements 1113

Soft-ferro core Induction coil PSD

Residual μr Faraday


B0 B μ,2ω V 2ω
(periodic in 2ω) induction
— 2ω
SATURATION
optional V̄
filter
Modulation flux
B ω B μ,ω compens. V ω
(periodic in ω)

Fig. 10 Fluxgate principle. The upper and lower halves, respectively, represent the transforma-
tions of the measured field B 0 and the modulation field B ω . After driving the ferromagnetic
core periodically into saturation, the modulation field is eliminated by compensation or frequency
filtering. Likewise, the constant measured field does not produce an induced voltage. However, the
flux it generates via the time-dependent residual permeability μr gives rise to an output voltage
that is proportional to the measured field

devices on a large industrial scale while avoiding the use of winding machinery
with moving parts. Due to their inability to detect constant fields, induction coils
rarely serve as stand-alone general-purpose sensors. They are mostly fabricated for
specific applications or to form part of a larger device. Simple pick-up coils are,
however, the standard monitoring tool for pulsed magnetic fields.
Fluxgate magnetometers [127, 139, 142] combine the rugged technology of
induction sensors with the capacity to measure constant fields. Figure 10 illustrates
their principle: a drive coil forces a ferromagnetic core periodically into saturation
thereby restricting its capability to respond to the external field. With respect to
the latter, the core exhibits a residual permeability that has maxima whenever the
drive field vanishes, i.e., at twice the drive frequency. Taking into account symmetric
distortions, the output of an induction coil that captures the total flux permeating the
core thus consists of odd harmonics caused by the drive field, whereas all even
harmonics are due to the external field.
The use of phase-sensitive detection techniques to recover the relevant second
harmonic from a fluxgate signal is relatively straightforward except that the
background of odd harmonics occupies most of the dynamic range and hence limits
the sensitivity. Advanced sensors are therefore designed such that the induction coil
does not capture any net flux from the drive field. As shown in Fig. 11, this can
be done by either magnetizing the core perpendicular to the induction coil axis
(orthogonal type) or using a double or ring core whose antiparallel flux contributions
cancel each other (parallel type). Still more sophisticated second-harmonic feedback
magnetometers make use of an additional compensation coil to operate as null-
detection devices.
The technical improvement of fluxgate magnetometers usually comes down to
a trade-off between the strict optimization of parameters determining the transfer
function in Table 3 and the control of secondary effects associated with these
optimizations. In complete analogy to induction coils, the addition of windings to
increase the effective area A of the sense coil is thus limited by the creation of
1114 O. Portugall et al.

Fig. 11 The basic parallel (left) and orthogonal (right) fluxgate configurations. Arrows indicate
the external field (red) and the magnetization of the core (yellow)

parasitic capacitances that ultimately affect the sensor response. Likewise, the shape
of the core influences not only the transfer function via the demagnetization factor
N but also the noise and modulation properties of the sensor. In the latter respect,
racetrack-shaped cores provide one of the best combinations of low noise and high
sensitivity [142]. Their layout resembles the parallel configuration in Fig. 11, except
that both cores are connected at the top and at the bottom so as to form a closed
magnetic circuit.
The core material enters the transfer function only via the residual permeability
μr and its time derivative, yet its choice is crucial for a variety of properties
including sensitivity, noise, linearity, field range, and power requirements. Materials
featuring a high permeability provide good sensitivity, but rapid magnetization
changes during operation tend to introduce noise associated with the domain
dynamics. The latter can be material-specific but ultimately also depends on the
technical quality of the core, in particular its homogeneity. Another important
material property is the saturation magnetization that determines not only the upper
limit of the measurable field range but, together with the coercivity, also the energy
consumption of the device. As fluxgate sensors are mostly optimized to provide best
sensitivity, the saturation field is usually kept small and the coercivity consequently
minimized. Suitable core materials should furthermore exhibit the least possible
magnetostriction to avoid vibrational perturbations during operation and have a high
resistivity to minimize eddy currents. While the criteria associated with different
applications may still motivate the choice of specific core materials, the majority of
fluxgate sensors nowadays make use of Co-based amorphous alloys or other metallic
glasses [142].
Being active devices, fluxgate magnetometers also depend on operating param-
eters. Square modulation generally assures the highest fluxgate output and hence
the best sensitivity. However, it also implies a relatively high energy consumption
depending on the core’s coercivity and saturation field. A possible work-around is
22 Magnetic Fields and Measurements 1115

to operate the fluxgate on a minor hysteresis loop, albeit at the price of sacrificing
part of the sensitivity [43]. Needless to say, the maximum modulation frequency
depends on the magnetization dynamics and eddy currents in the core.
Although their historical application range continues to be reduced by more
recent technologies, fluxgate sensors still represent the most appropriate choice for
measuring fields between roughly 10−9 and 10−12 T. The lower limit is associated
with their capacity to replace SQUIDs whose technology and operating conditions
are substantially more cumbersome, while the upper limit denotes their substitution
by even simpler magnetoresistive devices. While the technological optimization
of classical macroscopic fluxgate devices may be nearly exhausted, their minia-
turization using modern micro-fabrication techniques still shows some promise
[169]. Interestingly, the combination of deposition and etching in advanced MEMS
production techniques can replace coil winding which represents an undeniable
advantage for mass production.

Hall, Magnetoresistance (MR), and Magnetoimpedance (MI) Sensors


Magnetic fields can change the electric response of conductors by influencing the
drift of charges either directly via the Lorentz force or indirectly via their effect
on magnetic scattering centers. In conjunction with different boundary conditions
and measurement configurations, these primary field effects give rise to a rather
large number of charge transport phenomena that can be used in sensors. Figure 12
summarizes the principal, practically relevant, mechanisms.
Hall bars and Corbino discs, Fig. 13, are archetypical conductor geometries
that respectively constrain and short-circuit the lateral deviation of charges by the

Field effect Macroscopic setup Transport effect

Space charge
formation:
B0 Boundary Hall effect
Lorentz force:
conditions, V
charge deviation
I composites* Scattering on a
longer trajectory:
geom. MR, EMR*

Spin-orbit
I Direction
scattering:
control: I, M
AMR
Torque: magnetic
B0 V
scattering centers
Spin-dep. scatt.,
Magn. super-
& accumulation:
structures
GMR, TMR

Fig. 12 Hall and magnetoresistance sensor principles. Hall effect and geometrical MR are caused
by the Lorentz force in combination with suitable boundary conditions. EMR makes additional use
of composite structures. AMR, GMR, and TMR all refer to the intermediate effect of a magnetic
field on localized d states that subsequently act as scattering centers
1116 O. Portugall et al.

I
B
B B

Fig. 13 Hall bar (left) and Corbino disc (right). In the Hall bar, the deviation of charges by the
Lorentz force is limited by lateral boundaries and thus creates space charge regions on either side.
In a Corbino disc, the same effect produces a magnetic field-dependent circular current component
that prolongs the trajectory between the inner and outer disc radius and hence gives rise to MR

Lorentz force. In the first case, charges accumulate near the edges of a conducting
strip thus giving rise to a transverse electric field E that compensates the effect of the
Lorentz force, v × B. The Hall voltage V = Ew across a bar of width w therefore
provides a measure for the applied magnetic field. In Corbino discs, the situation
is different: here, the Lorentz force adds a circular component to the radial current
flow that prolongs the charge trajectory and hence increases the effective device
resistance. The so-called geometrical MR thus relies on the amount of scattering
albeit without influencing its origin. The same is true for the extraordinary MR
(EMR) that occurs in planar composite structures when the Lorentz force pushes
currents from low-resistance into high-resistance areas or vice versa [159, 72].
In order to generate significant effects, the Lorentz force has to produce a notable
curvature in the trajectory of charges between successive scattering events. Devices
are therefore based on high-mobility materials featuring long mean-free paths, while
their application is normally restricted to fields exceeding 10−5 T.
In magnetic materials, MR effects occur at substantially lower fields. Here,
the same, relatively localized, transition metal d-states that give rise to magnetic
properties also contribute to the scattering of delocalized states that govern the
charge transport. It is via these d-states that a magnetic field produces simultaneous
changes in the magnetization and MR whose phenomenological appearance can
be similar. The resistance changes nevertheless refer to microscopic scattering
processes and not just the flux-concentrating effect of the demagnetizing field.
The perturbative nature of scattering makes it generally difficult to pinpoint
microscopic mechanisms that govern a particular macroscopic behavior. Under
different conditions, similar mechanisms may even have opposite effects. Spin-flip
scattering thus accounts for the anisotropic MR (AMR) which is in principle a bulk
effect but tends to destroy the giant MR (GMR) that occurs in thin-film structures.
22 Magnetic Fields and Measurements 1117

AMR is a consequence of spin-orbit scattering between conduction band states


and magnetic d-states that depends on the angle between the linear momentum of
the former and the angular momentum or quantization axis of the latter [101]. On
a macroscopic level, this corresponds to the directions of the applied current and
the magnetization: if φ denotes the angle between both, and ρ and ρ⊥ are the
resistivities parallel (φ = 0) and perpendicular (φ = 90◦ ) to the magnetization axis,
then the resistivity for arbitrary angles is ρ(φ) = ρ⊥ +(ρ −ρ⊥ ) cos2 φ. AMR-based
sensors make use of the fact that a magnetic field tilts the magnetization relative
to its easy axis, thereby changing φ and hence ρ(φ). To provide best linearity
and sensitivity, devices are constructed so that the current flows at an angle of
45◦ with respect to the easy magnetization axis. The use of thin films essentially
constrains the possible directions of the magnetization as the shape anisotropy
strongly enhances the perpendicular demagnetization factor.
GMR refers to the coexistence of, ideally independent, currents of spin-up (↑)
and spin-down (↓) electrons that can only be scattered into unoccupied d-states
with the same spin [167]. In the absence of spin-flip scattering, the resistivity thus
includes a term ρs = ρ↑ ρ↓ /(ρ↑ + ρ↓ ) whose limits, 0 ≤ ρs ≤ ρ↑ /2 = ρ↓ /2,
are associated with ferromagnetic and antiferromagnetic configurations: in the first
case, the majority d ↑ -states are fully occupied which prevents scattering and hence
yields ρ↑ = 0, whereas in the second, both d ↑ - and d ↓ -states are partially occupied
giving rise to equal scattering and hence ρ↑ = ρ↓ . GMR sensors can exploit this
mechanism under two conditions: firstly, to avoid spin mixing, relevant dimensions
should not exceed the spin diffusion length, and secondly, the sensor must contain
sub-ensembles of ferromagnetic d-states whose relative orientation is sensitive to
a magnetic field in the desired range. Technically, this is achieved in thin-film
structures alternating non-magnetic with magnetic layers whose spin orientation is
antiparallel at B = 0 and gradually aligns at higher fields. The basic GMR effect is
observed in devices where the current flows in the layer plane (CIP = “current-in-
plane”). Advanced architectures make additional use of continuity conditions when
the current has to cross boundaries (CPP = “current-perpendicular-to-plane”): here,
the spin transfer across a non-magnetic layer is partially blocked and thus gives rise
to spin accumulation when the majority spins in the adjacent magnetic layers are
different. The effect is furthermore amplified in magnetic tunnel junctions (MTJ)
where the intermediate layer is insulating and the transfer occurs by tunneling
(TMR = tunnel MR). Further information on these architectures may be found in
 Chap. 31, “Magnetic Sensors” of this book.
For practical applications, common advantages of Hall and MR sensors are their
small size, typically in the μm range or less, their integrated circuit compatibility,
and, more generally, their suitability for mass production using modern thin-
film technology. A common disadvantage lies in their temperature sensitivity that
affects drift and offset stabilities as well as noise characteristics. Apart from
thermal Johnson noise related to their resistance and, inevitably, 1/f noise, devices
making use of semiconductors and ferromagnets may also be affected by shot and
magnetic noise. The use of ferromagnets in MR devices furthermore bears the
risk of irreversible temperature or field-induced magnetization changes. In some
1118 O. Portugall et al.

Table 4 Transfer function of Hall sensors in its usual form depending on the current density j ,
the Hall constant RH and the Hall bar width w, and an equivalent description in terms of the
mobility μH , the charge carrier concentration n, and the power density P /V
Type Transfer function
√ √
Hall effect U/Bex = j RH w = μH /ne P /V w

cases, devices are therefore equipped with small coils to re-magnetize relevant
components.
Their simple architecture and linearity makes Si-based Hall sensors an all-
round solution for measuring fields down to 10−3 T, a limit that can be extended
further to approximately 10−7 T with high-mobility materials such as InSb. The
principal factors that determine the theoretical performance of Hall sensors are
defined by the transfer function shown in Table 4, which depends on mobility, carrier
concentration, size, and the applied electrical power density. The optimization of
real devices is obviously more complex. For example, increasing the power density
not only raises the signal level but also affects the device temperature and stability.
Likewise, the choice of materials, suitable geometries, and advanced electronic
circuits for controlling, biasing, stabilizing, and readout may have an impact on
performance [175, 119, 124]. Ongoing research and development activities are
furthermore concerned with complex integrated devices such as miniature three-axis
sensors, making use of the self-positioning effect of micromachined strained-layer
structures [169].
MR devices can measure substantially lower fields than Hall sensors and gener-
ally dominate in applications between 10−3 and 10−9 T. Their intrinsic nonlinearity,
however, tends to restrict their operating range. Therefore, most MR sensors are
either taylored to fit specific applications and parameter ranges or constructed as
null-detection devices. A frequently stated performance criterion is their relative
resistivity change Δρ/ρ which can be misleading as it does not determine their
ultimate performance: AMR exhibits a Δρ/ρ of only 2–4% as opposed to values
exceeding 100% for GMR and 500% for TMR at room temperature. However, all
three MR types have comparable low-field detection limits around 10−9 T. The latter
criterion also explains why EMR has not achieved more practical importance despite
a Δρ/ρ approaching 106 %: here, the measured MR change refers to fields of the
order of 1 T [159]. Similar restrictions apply for intrinsic MR properties such as
colossal and “unstoppable” MR that have been observed in some materials [134, 3].
Minor strengths and weaknesses notwithstanding, there is no striking difference
between the overall performance of AMR, GMR, and TMR, and the choice of one
or another is usually motivated by application-specific requirements [176, 165].
AMR sensors are simple, cheap, and robust, and their use and technical devel-
opment continues despite the rapid evolution of GMR and TMR and the fact that
their physical properties were comprehensively analyzed a long time ago [101, 97].
To eliminate their resistance offset and temperature instability, AMR elements with
±45◦ current orientation are often combined in two differential pairs forming a
Wheatstone bridge [86]. A remaining disadvantage is their sensibility to orthogonal
22 Magnetic Fields and Measurements 1119

fields which should not exceed 10%. Notable recent developments of AMR sensors
include flexible devices [86] and μm-sized structures where the current flows
straight through a wide zigzag-shaped permalloy band that produces an alternating
±45◦ magnetization [30].
GMR and in particular TMR sensors are primarily used in information tech-
nology. As far as sensor applications are concerned, their integrated circuit com-
patibility and small size make GMR devices interesting for applications requiring
high speed and spatial resolution [167]. Other notable developments include GMR-
based integrated systems for 3D magnetic sensing or high-temperature applications
[84, 86]. TMR has all but replaced GMR in information technology and is also
advancing in sensor applications. Despite their inherent problem with shot noise
in the tunnel barrier, TMR-sensor prototypes reaching 10−10 T have been reported
[95]. A remaining disadvantage of TMR is that its functioning depends crucially on
the integrity of the insulating layer: pin holes destroy the device.
GMI (giant MI) [122] refers to magnetic field-induced changes of the driving
voltage that is necessary to maintain a radio-frequency (RF) current through a
ferromagnetic conductor. The mechanism can be easily understood on the basis of
Ampere’s circuital law: the RF current is associated with a magnetic field BRF ,
and both are subject to the same skin effect. The penetration of BRF , however,
is determined by the polarizability of the medium, i.e., its permeability that, in a
ferromagnet, depends on the applied magnetic field B0 . Changing B0 thus tunes the
effective conductor cross section and hence the RF impedance. The effect becomes
“giant” in the vicinity of characteristic time constants of the medium that may be
associated with, for example, collective excitations, relaxation mechanisms, or the
domain dynamics.
GMI sensors use rather simple probe heads that essentially consist of a small
length of amorphous wire but inevitably require RF circuitry for their operation.
Industrial applications exist [74], and substantial research effort is dedicated to
the development of new GMI materials [122]. On the other hand, the precise
mechanisms giving rise to GMI are not yet fully understood, and theoretical
predictions promise much larger effects than those actually observed. GMI thus
remains an emerging technology, albeit with interesting perspectives: its response
ρ/(ρB) can reach 5000%/mT as opposed to 10%/mT for GMR devices [122].

Superconducting Quantum Interference Devices (SQUIDs)


One of the most remarkable features of superconductivity is the coherence of the
wave function of superconducting carriers. This allows the observation and exploita-
tion of phase-dependent phenomena on a macroscopic scale [168]. Superconducting
quantum interference devices (SQUIDs) are still the most sensitive devices for
measuring changes in magnetic field and magnetic moment. They are based on two
phenomena from the physics of superconductors – the Josephson effect and flux
quantization [87, 32].
A Josephson junction (J J ) consists of two superconductors separated by a
weak link. The weak link is a thin layer of insulator, or normal metal, or a
tiny superconducting bridge, which allows the wave functions of the neighboring
1120 O. Portugall et al.

Fig. 14 Josephson junction, with the corresponding I − V characteristic

superconductors to overlap but gives rise to a phase drop. If the phase drop is
nonzero, the junction is penetrated by a persistent current, not exceeding, however,
a critical value IC (Fig. 14).
Magnetic flux penetrating a closed superconducting circuit is quantized and can
only take values of integer numbers of flux quanta:

h
Φ = nΦ0 , where Φ0 ≡ ≈ 2.068 · 10−15 Wb (2)
2e

If the circuit is interrupted by a Josephson junction, flux quantization has to be


generalized to take into account the phase drop:

2π n = Δφ + 2π Φ/Φ0 (3)

DC SQUIDs. Let us now consider a superconducting circuit with two J J s with


the critical current IC . Let us assume that a current I is sent through the circuit
(Fig. 15a). Let us further assume that this current is ramped from zero to a value
exceeding 2IC and back in order to probe the critical current of the whole device
SQU I D SQU I D
IC . If the circuit is symmetric, then, in the absence of external flux, IC =
2IC . If now the external flux Φ < Φ0 /2 is applied to the circuit, a screening current
j sets in. Then the current in the left-hand branch of the SQUID will be I /2 + j and
in the right-hand one I /2 − j . Thus in the right-hand branch, the maximum possible
current will be reduced, and in the left-hand branch, it will be maintained, as it is
SQU I D
limited by IC of the J J . Thus, the net critical current IC will decrease.
As the external flux exceeds one half of the flux quantum, due to the kinetic
energy Lj 2 /2, it is more favorable for the system to let one flux quantum penetrate
the circuit with the opposite sign, and reverse the screening current j (Fig. 15b).
Further increase of the external flux will now reduce |j |, and ICSQU I D will grow
SQU I D
until the external flux reaches Φ0 . Here the IC will be restored to its original
value (Fig. 15c). As the external flux is ramped, the I −V curve will evolve between
22 Magnetic Fields and Measurements 1121

Fig. 15 DC-SQUID operation: (a) Schematic circuit. (b) Screening current j (Φ). (c) SQUID
critical current ICSQU I D (Φ). (d) I − V curves for integer and half-integer flux quanta. (e) V (Φ)

the two extrema (Fig. 15d). If we now set the bias current in between, then the
corresponding voltage will oscillate with the period of one flux quantum (Fig. 15e).
RF SQUID. The setup with two J J s described above is called DC-squid, as it
is biased by a DC current. Opposed to this, an RF-SQUID is a superconducting
loop interrupted by only one J J and inductively coupled to an RLC “tank” circuit
(Fig. 16a). The tank is fed with a radio-frequency (rf) current at frequencies close
to the resonance of the circuit. Typically, the SQUID is designed so that it is able
to compensate about one flux quantum before the critical current IC is reached.
Let us assume that initially the SQUID contains an integer number of flux quanta
nΦ0 . If we now apply the rf-flux by ramping Irf of the tank, the SQUID has to
compensate it to maintain the flux in the circuit. As long as the screening current is
lower than IC , the corresponding voltage Urf will grow linearly as the SQUID is
in a fully superconducting state and does not affect the tank impedance. Once the
screening current reaches IC , the SQUID will change the state to n+1 or n−1. As
this transition takes place, the energy is drained from the tank, and it takes some time
(several RF periods) to recover the energy. As the energy is restored, the transition
takes place again. If Irf is increased, the tank energy recovers faster, so the flux
transitions in the SQUID are more frequent. Averaged over many flux transitions,
1122 O. Portugall et al.

Fig. 16 (a) RF SQUID inductively coupled to a “tank.” (b) V − I curves of the tank at integer
and half-integer flux quanta. (c) Tank voltage VT (Φ)

this dissipative process shows up as a plateau on the I − V characteristic of the


tank (Fig. 16b). If we increase the external flux Φ, the transition will occur at lower
values of Irf , and the dissipative plateau will shift down until it reaches its minimum
at Φ = 1/2Φ0 . Therefore, if the bias RF-current is set so that the corresponding
plateaus overlap, the voltage drop over the tank will depend on the applied flux as
shown in Fig. 16c.
Both DC- and RF-SQUID work as flux-to-voltage converters within a fraction of
a flux quanta [24, 25]. Their sensitivity is limited by two types of noise – thermal
and 1/f , also called “flicker” noise.
Real devices usually include a feedback, so that the flux in the SQUID is
maintained in the vicinity of the optimal working point (null-detector configuration).
For an RF-SQUID, the tank voltage is first rectified, and the obtained quasi-DC
voltage is applied back to the SQUID (phase locking). Along with the low-frequency
flux modulation, phase locking helps diminish the 1/f noise.
Eventually, the practical signal-to-noise ratio is defined by the temperature and
environmental noise. Nowadays, both types of SQUIDs are capable to reach similar
sensitivity values of about 10−6 Φ0 , or 10−21 Wb. This sensitivity margin allows
for very high spatial resolution as well. The prerequisites for using SQUIDs are a
cryogenic environment and external noise shielding.

Nuclear Magnetic Resonance (NMR) Sensors


Magnetic resonance-based methods provide nowadays the most accurate mea-
surements of static and homogeneous magnetic fields. In favorable conditions,
precisions of 0.1 ppm and better with direct traceability to primary standards within
1 ppm accuracy are obtained. This is due to the fact that these methods measure the
field strength by a frequency measurement, i.e., a simple counting experiment that
can be realized with very high precision.
Magnetic resonance methods use the interaction of a localized magnetic moment
with an external magnetic field. In the presence of a magnetic field B0 , the magnetic
moment m of a nucleus with nuclear spin I > 0 or an unpaired electron carrying
22 Magnetic Fields and Measurements 1123

a spin S in paramagnetic molecules and solids exhibits a Zeeman splitting of its


discrete energy states. This splitting is proportional to the applied magnetic field.
Taking the simplest case of S or I equal to 1/2, a magnetic dipole transition occurs
at a characteristic resonance frequency, the Larmor frequency ωn that is proportional
to the applied magnetic field B0 . This phenomenon is called nuclear magnetic
resonance (NMR) for the case of nuclear spin states and electron paramagnetic
resonance (EPR) for the case of electronic spin states. Using an appropriate
detection setup, a time-dependent signal can be obtained. Its Fourier transform
allows the determination of ωn and thus B0 .
Signal properties and precision. In order to obtain quantitative information
about the signal properties and the precision of the method, a brief description
of the NMR/EPR method will be presented. A comprehensive review of magnetic
resonance techniques for field measurements can be found in [137]. As NMR and
EPR are based on the same formalism, we will restrict us in the following to the
NMR case, which is the more widely used method. This is due to the fact that EPR
methods require paramagnetic compounds of uncoupled isotropic electron spins,
which involve radicals. These compounds are very often found to be chemically
unstable or toxic.
The energy of a nuclear spin I with a magnetic moment mN = γn I in
an external magnetic field B0 is given by U = −mN · B0 ; γn denotes the
gyromagnetic ratio, which is a generic property of the observed nucleus that is
defined in tables. For shielded protons, 1 H, in water, H2 O, at 298.15 K γn /(2π )
is 42.57638507(53) MHz/T using the 2014 CODATA recommendation [106]. For
other nuclei, the gyromagnetic ratios can be obtained using tables [64]. It is
important to note that the chemical environment of the nucleus and the temperature
affect this value. Therefore, care must be taken to use appropriate values of
gyromagnetic ratios, particular for high-precision magnetic field measurements.
Taking the simplest case of I =1/2, as for protons 1 H, the magnetic field generates
quantized energy states denoted by MI = ±1/2 that are separated by h̄ω with
ω = γn B0 . This energy contribution is called Zeeman energy (see Fig. 17).
The phenomenon of NMR can be understood by the following semiclassical
consideration. A nuclear spin I with a magnetic moment mN = γn I in a magnetic
field B0 is subject to a torque Γ = mN × B. As a consequence of the equation of
motion of angular momentum in the presence of a torque, a precession of the nuclear
spin occurs around the axis of B0 with the Larmor frequency ωn = γn B0 . For a
detailed understanding of magnetic resonance phenomena, the reader is referred to
standard textbooks [1, 157].
The Larmor frequency varies linearly with the field, and its measurement is
the basis of all field measurement techniques by magnetic resonance. The field
measurement is hence transformed in a frequency measurement that can be made
with a relative precision of 10−12 . Using standard radio-frequency (RF) technology,
the application of NMR field measurement techniques ranges from the MHz to
the GHz range covering field ranges from 10−2 T to 102 T. Lower fields can be
accessed using EPR as the gyromagnetic ratio of the electron is almost three orders
1124 O. Portugall et al.

Fig. 17 Quantum (left) and classical (right) NMR principle. Left: Zeeman energy of discrete
energy levels of a spin-1/2 system in the presence of a magnetic field B0 . At the Larmor
frequency ω = γ B0 , a resonant absorption occurs due to a magnetic dipole transition. Right:
The magnetic moment associated with angular momentum precesses in an external magnetic field
with the Larmor frequency. This precession can be excited and detected by a pick-up coil oriented
perpendicular to the external field

of magnitude higher, 28.025 GHz/T [37], or using hyperpolarization techniques


leading to accuracies in the fT region [59].
An NMR-based field measurement setup typically consists of a mm-size NMR
sample that is placed into the magnetic field of interest. The nuclear magnetization
will serve as the field detector. The sample is surrounded by a coil, whose axis is
aligned perpendicular to the main field and that is part of a tunable resonant RF
circuit. Other elements are a precise tunable RF source as well as RF components
for excitation and detection [54].
As NMR samples consist of an ensemble of uncorrelated nuclear spins, the
primary field sensor is the average macroscopic nuclear magnetization. In thermal
equilibrium, this magnetization is aligned along the external field B0 , and no
precession occurs. In standard conditions for NMR field measurements, the sample
is furthermore at room temperature. Hence, the nuclear spin system is paramagnetic,
and the equilibrium magnetization is proportional to B0 . In order to observe a
precessing average magnetization, a non-equilibrium state has to be generated. This
is achieved by an external continuous or pulsed radio-frequency field B1 (t) =
B1 cos(ωn t) that is tuned to the Larmor frequency ωn and oriented perpendicular
to the main magnetic field B0 . B1 (t). It is generated by the excitation coil. After
application of B1 (t), a nonzero precessing magnetization occurs that can be detected
by its interaction with the excitation coil that now acts as pick-up coil. After
amplification, the voltage signal of the pick-up coil is recorded and used for the
determination of the magnetic field (see Fig. 18).
The factors that limit the sensitivity and accuracy of an NMR experiment are
mainly the strength of the signal caused by the precessing magnetization, quantified
by its signal-to-noise ratio, and its coherence time that limits the spectral resolution
and hence the accuracy of the method. In the case of a detection by a pick-up
coil after an excitation pulse, the signal is given by the induced voltage. It is
proportional to B02 , as the pick-up signal is proportional to the time derivative of the
22 Magnetic Fields and Measurements 1125

Nuclear spins Pick-up Detection

Zeeman
splitting PSD (V )

induction
Faraday
at ωn
B0 M (t) dM
dt
torque, or
precession FFT (V )
ωn = γB0

Fig. 18 NMR detection principle

magnetization (∝ B0 ), which precesses at the Larmor frequency (∝ B0 ). Adding


9/4
the influence of noise, the signal-to-noise ratio is finally found proportional to B0
[76]. Other factors that affect the signal are the natural abundance n of the isotope
carrying a nuclear spin and the gyromagnetic ration, γn , of the nucleus involved.
At fixed magnetic field B0 , the signal-to-noise ratio is found proportional to nγn3 .
Therefore, it is convenient to use NMR active nuclei with high abundance and large
nuclear gyromagnetic ratios, such as 1 H. State-of-the-art NMR instruments provide
signal-to-noise ratios of the order of 1000:1 for 1 H NMR of a 1 mm3 sample at
frequencies of 100 MHz, corresponding to B0 of 2.3 T. The second factor, the
coherence time, is influenced by nuclear relaxation processes as well as spatial
inhomogeneities and temporal instabilities of the external magnetic field. All these
effects can be modelled by a time constant T2∗ that describes the exponential decay
of the signal after the excitation pulse. For liquids at room temperature, upper
limits of T2∗ are caused by nuclear relaxation processes and are of the order of
1 s with record values in gases for T2∗ of 60 h [59]. In frequency domain 1/T2∗
is proportional to the spectral linewidth that provides a measure for the resolution
and hence the accuracy of the technique. Using the value of 1 s for liquids cited
above, the resolution becomes better than 1 ppm for frequencies above 1 MHz or B0
above 20 mT.
Various setups and techniques for NMR-based field measurements are used. The
first two methods are continuous wave techniques based on measuring either the
variation of the Q-factor or the impedance of a tuned resonance circuit using closed-
loop circuits. In these cases, the frequency of the detection circuit is latched on
the Larmor frequency ωn of the nucleus. The third method is a transient method.
The nuclei are excited with a short resonance pulse that flips the equilibrium
magnetization into a direction perpendicular to the external magnetic field. The free
precession of the nuclei are then detected by a pick-up-coil: free induction decay
(FID). Its functional form is a damped oscillation at ωn decaying with T2∗ . All three
methods are commercially available [137].
Special environments. The ideal field conditions for NMR are homogeneous and
time-stable fields, and the above given accuracy and precision is valid for these
conditions. The presence of spatial variation or time variation hence strongly affects
the performance of the technique that will be discussed in the following.
1126 O. Portugall et al.

Spatial variation of external magnetic field: A spatial variation of the external


magnetic field B0 generates a spatial variation of the Larmor frequency ωn over
the sample volume. This induces an additional time decay of the NMR, i.e., T2∗
will become shorter. Hence, an additional line broadening occurs that reduces
the resolution of the technique. In order to avoid this effect, one can use smaller
samples with the drawback of a reduced signal-to-noise ratio. For samples with
1 mm dimension, gradients of the order of 10 to 100 ppm/mm are tolerable without
significantly reducing the accuracy of the method. In this context, it should also be
noted that the Larmor frequency always measures the modulus of the magnetic field,
independently of its direction. However, the field direction affects the amplitude of
the signal.
Since NMR field measurement methods are local probes with a spatial resolution
that is given by the sample volume, they can be used to precisely map spatial profiles
of magnets. This has become an important tool for validation of theoretical magnetic
field profiles. For this purpose, two methods exist. For the first method, a single
sample is put at various positions in a magnetic field. At each position, the value of
the magnetic field is recorded that finally provides a field map [172]. The drawback
of this method is that the field values are not recorded simultaneously so that time
variations of the magnetic field can influence the result. This drawback is partially
eliminated by the second method, which uses an array of NMR probes that are
located at different positions and fast readout electronics [23, 137].
Time variation of external magnetic field: Time variations of external magnetic
field can generate various effects on the field measurements by NMR. Very slow
time drifts occurring on timescales much longer than the nuclear spin relaxations
times can be directly tracked by the NMR. This also allows the construction of
field stabilization units (field-frequency locking) that are used in NMR instruments
for spectroscopy [126, 77]. However, temporal variations occurring on timescales
of the same order as the intrinsic nuclear spin relaxation times strongly reduce
the resolution of NMR-based field measurement techniques, since they generate
additional line broadening that can even completely suppress the NMR signal.
Very fast time variations are averaged out. However, they can also generate nuclear
relaxation processes.
Tracking range of NMR method: Although NMR can be used to detect a broad
range of magnetic fields from 10−2 T to 102 T with high precision, it becomes
difficult to track time evolution of magnetic fields over that range. Therefore, the
user has to provide a coarse value of B. The NMR field measurement then provides
the desired precise value as offset from that coarse field value within a given
tracking range. Standard NMR field measurement setups that are based on tuned
radio-frequency circuits have a tracking range that is limited by the Q-factor of
the radio-frequency circuit. With Q-factors typically varying from 50 to 1000, the
corresponding tracking ranges are between 2% and 0.1%.
Dynamic range of NMR method: As nuclear spin dynamics is instantaneously
influenced by the present value of B0 , it is possible to adapt the time window of
NMR-based field measurements. This is particularly interesting for absolute field
measurements at very high magnetic fields of fluctuating resistive magnets or for the
22 Magnetic Fields and Measurements 1127

tracking of absolute values in pulsed magnetic field experiments. Taking 1 H NMR


at 23.5 T, which corresponds to a Larmor frequency of 1 GHz, 1 ppm resolution
corresponds to a characteristic FID time constant of ≈ 1 ms. This means that one
can allow temporal field variations and still maintain a precision at the ppm level.
This has been demonstrated in pulsed magnetic fields up to 62 T [102].
Instumentation and cost. NMR-based field measurements always require a
complete experimental setup comprising a tunable NMR probe, high-frequency
electronics, and a highly stable and phase-coherent radio-frequency synthesizer,
which is the most expensive element of the setup. For standard applications
such as highly homogeneous and time-stable superconducting magnets with room-
temperature bore operating in the 1–10 T range, commercial instruments for field
mapping are available providing accuracies of 1 ppm or better [137].
NMR can also be used for the precise field mapping in nonstandard applications
involving special magnet geometries, high magnetic fields, time-varying magnetic
fields, or low temperatures. However, tailored NMR probes and field measurement
strategies need to be developed in this case.

Nonresonant Electromagnetic Probing: Faraday Rotators


Cyclotron motion and Larmor precession affect the polarizability of a medium
and hence the propagation of electromagnetic waves. The latter can generally be
decomposed into 2 normal modes, i.e., linearly independent solutions of the wave
equation whose polarizations are conserved despite the interaction. Absorption
occurs when the electric or magnetic field vector of one of the modes is constantly
in phase with the motion of free charges or magnetic moments and a net transfer
of energy takes place. This gives rise to resonant effects such as NMR, EPR,
or cyclotron resonance. In the absence of absorption, the normal modes still
exhibit different phase velocities. Linear combinations of both thus undergo a
field-dependent polarization change called circular magnetic birefringence that
produces magneto-optic phenomena such as the Faraday and Kerr effects (see
also  Chap. 10, “Magneto-optics and Laser-Induced Dynamics of Metallic Thin
Films”).
Figure 19 illustrates the principle of a Faraday rotator: linearly polarized
radiation propagating in the direction of an applied magnetic field is composed of

Rotator medium Polarizer Detector


B0

hν Cyclotron motion
Projection

Intensity

Larmor precession
hν hν hν V
Circular magnetic
hν birefringence

Fig. 19 Magneto-optical detection principle (Faraday and Kerr effect). Red arrows indicate the
polarization at different stages of the measurement process
1128 O. Portugall et al.

Table 5 Transfer function of a Faraday rotator with ϑ as the rotation angle, V the Verdet constant,
l the length of the optical path, I0 the incident intensity, and ΔI the intensity change. In the
second formula, sin ϑ ≈ ϑ has been assumed. With standard optical detectors, the intensity change
translates directly into a voltage
Type Transfer function Condition
Rotation angle ϑ/B0 = V l Below magnetic saturation
Intensity change (I0 /2 + ΔI )/B0 = I0 V l 45◦ -polarizer, B0  (V l)−1

left- and right-hand circularly polarized normal modes. Depending on whether their
polarization vector revolves with or against the circular motion of free charges or
magnetic moments, the modes are subject to different phase shifts that ultimately
rotate the original linear polarization. To identify the rotation angle ϑ ∼ B0 ,
Table 5, an optical detector measures the relative radiation intensity transmitted
through a polarizer. This is proportional to ∼ cos2 (ϑ + ϑ0 ) where ϑ0 is the angle
of the polarizer relative to the polarization of the incident radiation. The polarizer
orientation determines how much intensity is transmitted at B0 = 0 and how it is
affected when the field is raised. In the particular case of ϑ0 = 45◦ , the change
of transmission associated with a sufficiently small magnetic field evolves linearly
around half the value of the incident intensity and depends on the direction of the
field.
Faraday rotation is material and wavelength dependent as characterized by
the Verdet constant. The latter becomes large in the vicinity of absorption lines
where the refractive index exhibits strong dispersion. Absorption must, however, be
avoided in order to guarantee that intensity changes are solely due to polarization.
Popular materials for Faraday rotators include terbium gallium garnet (TGG,
Tb3 Ga5 O12 ) featuring a large Verdet constant of −134 rad/T·m at 632.8 nm wave-
length [177]. Like all magnetic materials, TGG is subject to saturation and hence
becomes nonlinear at higher fields [178]. Typical Verdet constants of dielectric
materials are substantially smaller but, for all practical purposes, constant. CdS has
been reported to possess a particularly high Verdet constant around 125 rad/T·m at
632.8 nm [130] while evolving linearly up to about 500 T [38]. This corresponds to
one full rotation every 50 T over a length of 1 mm.
Faraday rotators can have extremely fast response times as the process itself is
only limited by the speed of light. Sensitivities may vary substantially depending
on the complexity of the setup [43]. The lower field limit in Table 2 represents an
average in this respect.
Modern fiber-optic technology has permitted the production of compact Faraday
rotation devices by integrating both polarizing and magnetically birefringent ele-
ments in a single fiber [166]. The latter are obtained by local terbium doping. The
practical importance of Faraday rotators as magnetic sensors remains nevertheless
limited compared to mass products such as Hall and MR devices. A notable
exception are applications where electromagnetic compatibility or remote sensing
over large distances is required. The frequency bandwidth, resolution, and noise
22 Magnetic Fields and Measurements 1129

characteristics of Faraday rotators are generally determined by intrinsic properties


of the optical detector.

Mechanical Sensors: MEMS-Based Magnetic Force Meters and


Magnetoelectric Laminates
The Lorentz force acting on a current filament, the torque experienced by a
bar magnet, and magnetostriction are the basic mechanical phenomena that can
be exploited for sensing applications. As illustrated in Fig. 20, the respective
magnetometers are generally two-stage devices where the initial mechanical effect
is subsequently converted into an electrical signal. Modern deposition and MEMS
fabrication techniques (micro-electro-mechanical systems) permit the integration of
both functionalities in a single miniaturized chip or even inside composite materials.
Most, but not all, mechanical devices make use of flexible structures such
as cantilevers, trampolines, xylophones, or torsion bars to create measurable
displacements. The resonance of the structure determines the upper bandlimit of
the detector. It can also be used to suppress noise and to increase the deflection
amplitude and hence the detector responsivity if modulation techniques are applied
and the device is operated at resonance frequency. Small current filaments on
xylophone or cantilever structures exploiting the Lorentz force are most suitable in
this respect [69, 169]: the current can be easily modulated and, moreover, adjusted
to provide optimum deflection for a given field range [88]. Tiny bar magnets on
perpendicular torsion wires [189, 35], magnetostrictive films or magnetoelectric
laminates on cantilevers [98] and cantilever structures made from magnetic
materials [22] provide less flexibility and dynamic range but have the advantage of
being completely autonomous as they draw power directly from the external field.
Their response is, however, intrinsically nonlinear. Bias and modulation fields are
commonly employed to overcome this problem, albeit at the expense of a larger size.

Mechanical effect Read-out options PSD

Opt. deflection
I Lorentz
Demodulation (optional)

force
Opt. diffraction
B0
Interferometric
B op Compass ε V
Capacitive

Piezoelectric
Magneto-
B op striction
Tunneling

Fig. 20 Mechanical sensing principles in MEMS. Any mechanical sensing principle can in
principle be combined with any readout method. B op stands for an optional bias or modulation
field. The phase-sensitive detection step (PSD) is also optional
1130 O. Portugall et al.

While bias fields can still be implemented with permanent magnets, modulating
the field requires electromagnets which also reduces the sensor’s autonomy. On the
other hand, mechanical sensors featuring tunable bias magnets can also make use of
feedback loops to operate at constant deflection.
Mechanical deflection can be measured with a variety of methods whose
suitability depends on the required sensitivity and the practical context [69].
Detecting the actual torsion or deflection with a deposited layer of piezoresistive
material represents the most direct approach. Other methods refer to the tilting
of a lever or the displacement of a platform with respect to a fixed reference.
Electric measurement techniques thus make use of planar capacitors in a resonant
LC-circuit whose electrodes are respectively attached to the flexible structure and
a static support. Alternatively, the tunneling current through tips mounted in the
same way can be measured. A variety of optical methods refers to the reflection,
diffraction, and interference of light beams. In the first two cases, position-sensitive
detectors determine the deviation of a beam by either a mirror or a grating. Here,
the detector arrangement typically consists of two receptors capturing different parts
of a comparably large diffuse light spot. Interferometric measurements make use
of mirrors forming a Fabry-Perot cavity between the flexible structure and a static
reference.
Mechanical sensing without moving parts is possible with flat, typically
centimeter-sized, sheets of magnetoelectric composites that consist of alternating
layers of magnetostrictive and piezoelectric materials [192, 179]. The mechanical
coupling transmits field-induced strain from the magnetostrictive to the adjacent
piezoelectric layers where it produces a voltage that can be measured with a
simple two-terminal setup. Magnetoelectric sensors thus also have the advantage
of zero power requirements. Some single-phase materials exhibit magnetoelectric
properties albeit at a level that is too weak for sensor applications.
Compared to other sensor types, mechanical magnetometers are an emerging
technology based on either the advent of MEMS fabrication techniques or recent
developments in the area of magnetoelectric laminates. Lorentz force-based devices
are slightly less sensitive (cf., Table 2) but have the advantage of being linear,
easy to modulate, and capable of covering a relatively large field range. The use
of magnetoelectric composites, on the other hand, has the potential to provide √better
sensitivities over a limited field and frequency range. Noise levels of 10 pT/ Hz or
less have been reported for both stand-alone laminates and thin films on cantilever
structures which make these devices a likely candidate for biomedical applications
[83]. MEMS devices and magnetoelectric laminates are likely competitors for
fluxgate magnetometers that dominate the field range between 1 pT and 1 nT. Their
frequency bandwidth is, however, limited.

Bulk Magnetic Measurements

Magnetic objects generate or alter magnetic fields in their environment. In the


following, we discuss magnetic sensor arrangements that are used to measure the
22 Magnetic Fields and Measurements 1131

effect of these fields in order to gain insight into the magnetic (spin) and electronic
(charge) properties of the object.

Induction Magnetometers
In Sect. “Induction and Fluxgate Sensors”, the induction method for measuring
on Faraday’s law, V = −∂Φ/∂t, where
magnetic fields was introduced. It is based
the magnetic flux Φ is defined as Φ = B·dS. In the presence of a magnetic
object, B = μ0 μr H, or, otherwise, B = μ0 (H + M). Here, μr is the relative
magnetic permeability, and M is the magnetization of the material, defined as the
magnetic moment of a unit volume of the material: M = dm/dV; m is the net
magnetic moment of the sample originating from atomic dipole moments.
Generally, the magnetic permeability is a complex second-rank tensor: μ =
μ0 μr = μ + iμ . In order to define it, all the components of magnetic moment
vector should be measured for a given field direction. Appropriate devices called
“vector magnetometers” exist, yet, for most purposes, it is enough to measure
the magnetic moment along the field direction. This covers both isotropic and
anisotropic cases, when the permeability tensor is diagonal. Thus, scalar magne-
tometers will be assumed further in the text unless explicitly specified otherwise.
Magnetic flux then can be split into two parts:

Φ = ΦH + Φ M = (μ0 H + M) · dS (4)
S

By induction-type magnetometer, we understand a device determining the


magnetic moment or magnetization of the sample by integrating the voltage induced
by a change of ΦM while the contribution of ∂Φ/∂t is rejected or separated.
In order to cancel the contribution of the time-dependent magnetic field, com-
pensated pick-up coil systems, also called gradiometers, are commonly used in
induction-type magnetometer devices. Figure 21 presents several possible arrange-
ments. All the types presented (except the single coil) are insensitive to a spatially
uniform time-varying magnetic field. First-order gradiometers (types (b) and (d))
also cancel linear field gradients H (t) = H0 + ∂H ∂r dr along the corresponding
direction. Second-order gradiometers additionally cancel terms up to second-order
∂2H
∂r 2
dr 2 . A coaxial gradiometer is able to compensate even higher-order gradients.
Extraction magnetometer: The sample is first placed in the measurement coil,
which is therefore penetrated by the total flux Φ = ΦH + ΦM , associated with the
applied magnetic field and the sample magnetic moment m. When the sample is
extracted from the coil, ΦM vanishes and leaves only the field contribution ΦH . In a
cylindrical coil whose axis is oriented along the z-direction this flux change induces
a voltage whose integral is proportional to the respective component of the magnetic
moment of the sample, i.e.,


mz ∝ Vind (t)dt (5)
1132 O. Portugall et al.

Fig. 21 Possible pick-up coil arrangements: (a) single coil, also called ‘magnetometer’, (b) first-
order gradiometer, (c) second-order gradiometer, (d) planar first-order gradiometer, (e) coaxial
gradiometer

Fig. 22 Extraction magnetometer: sensitivity profile and schematic drawing

The method described above is used in practice, yet, it is subject to noise due
to vibrations of the measurement coil and variations of the external magnetic field.
To cancel those spurious contributions, a pair of identical coils, connected in series
with opposite polarity (first-order gradiometer), is used. The sample is moved from
one coil to another, and the resulting integral is roughly twice the value of the
single-coil case. The coils are designed to provide flat parts on the sensitivity profile
in order to reduce dependencies on the positioning accuracy (Fig. 22). For certain
coil geometries, the sensitivity profile S(z) and the calibration constant C can be
calculated [193]. In practice, the whole device is calibrated with a known sample.
The particular time dependence of the sample movement is not important. There
are only technical limitations due to the speed of extraction: the movement should
be fast enough to provide an induced voltage in a comfortable range, and not too fast
to avoid disturbance of the integration. This allows for various types of drivers for
transporting the sample: electromagnet, stepper motors, or pneumatic pistons. An
extraction magnetometer can be conveniently built as a vector magnetometer [41],
22 Magnetic Fields and Measurements 1133

in which case, three orthogonal pairs of pick-up coils are used. Further improvement
can be achieved by averaging multiple extractions. The sensitivity can reach 10−8
Am2 [131].
Vibrating sample magnetometer: The vibrating sample magnetometer (VSM),
also commonly named Foner magnetometer after its inventor [51], utilizes a similar
principle of displacing the sample with respect to the measurement coils. However,
instead of taking the full integral of the sample flux, the sample is vibrating between
the coils in a time-harmonic manner:

z = A sin(ωt), (6)

where A is the amplitude of the sample vibrations and ω is the frequency.


If the response of the coil is linear, the induced voltage is then given by

Vind = Cmz ωA cos(ωt), (7)

where mz is the projection of the sample moment on the coil axis and C is the
calibration constant.
The vibration frequency ω appears as a factor in the effective sensitivity.
However, as considerable masses are involved in the motion, the frequency typically
lies within the range of 30–300 Hz, and the corresponding displacement amplitudes
vary from a few millimeters to a fraction of a millimeter. Phase-sensitive detection
(PSD) procedures are applied in order to convert the time-dependent induced voltage
to a quasi-DC signal proportional to mz . The oscillatory motion of the sample is
tracked by a dedicated sensor in order to feed the PSD with a proper reference
signal: it can be a small magnet attached to the driving rod, inducing a voltage in a
coil placed nearby.
The pick-up coil system is optimized to ensure a linear sensitivity region fully
covering the vibration amplitude [18]. It can be realized in both axial [75] and
transverse [51] geometry. In the former case, a first-order gradiometer is typically
used as pick-up coil. In case of transverse magnetic fields, the pick-up coil system
consists of a coplanar pair of flat gradiometers with the coil areas normal to the
magnetic field (see Fig. 23).

Fig. 23 Vibrating sample magnetometer


1134 O. Portugall et al.

Fig. 24 AC susceptibility measurement setup

In most cases, the pick-up coil system is fixed with respect to the environment,
and the sample is moved. Alternatively, it is possible to fix the sample and move the
sensing coils. This is called then vibrating coil magnetometer, or VCM [158]. It is
more sensitive to inhomogeneities of the external field but allows measurement of
mechanically unstable samples such as loose powders and liquids. The sensitivity
of a VSM can reach 10−9 Am2 .
AC susceptibility measurements: The magnetization M can be defined as dM =
χ dH. The susceptibility χ thus is given as dM/dH. It can be measured in practice
by applying an alternating magnetic field (AC field): H = H0 sin ωt that should
be small enough to provide a linear response ∂M/∂H . The sample is located in a
pick-up coil that is part of a first- or second-order gradiometer. The entire system
is surrounded by an excitation coil generating the AC magnetic field (Fig. 24). In a
perfectly compensated pick-up coil system, the induced voltage will then be given
by:

∂Φm ∂M
Vind = − =C = Cωχm H0 cos(ωt + φ) = CωH0 (χ  cos ωt + χ  sin ωt)
∂t ∂t
(8)
A phase shift φ or, otherwise, nonzero parameter χ  contains information on
the relaxation time of the spin system of the sample, as well as on dissipative
processes such as eddy currents or domain wall motion. Measurements of Vind are
typically performed with a lock-in amplifier (i.e., PSD), yielding both χ  and χ 
simultaneously. The AC field can be superimposed on an external DC magnetic
field, thus providing the field-dependent susceptibility χ = χ (H ). Integrating the
latter, the whole magnetization curve can be reconstructed. The AC susceptibility
technique is rather easy to implement in a compact size, and it is suitable for an
extremely wide temperature range of 1 mK up to 1000 K. The sensitivity of the
technique reaches 10−11 Am2 [190, 131].
Pulsed-field magnetometer: As described in the magnet section, magnetic fields
above 45 Tesla can only be achieved by pulsed magnets with typical durations of
10−6 –1 s. This is too short for a sample movement, and sometimes not even long
enough for AC-susceptibility measurements. Also, the requirements for compen-
sating gradiometers are extremely high. Unlike their DC counterparts, where the
22 Magnetic Fields and Measurements 1135

Fig. 25 Pulsed-field magnetometer

external magnetic field is either stabilized or ramped rather slowly, pulsed-field


magnetometers are exposed to enormous induced voltages as a consequence of the
amplitude and sweep-rate of the external field.
Magnetization measurements in pulsed fields simply exploit the fast ramping
as the excitation: the magnetic moment of the sample varies as the magnetic field
changes. This process is rather fast and induces a significant voltage Vm (t) dm/dt
in a coil surrounding the sample [4] (Fig. 25) that is superimposed by another
contribution from the external field VH (t) dH /dt . A second coil distant from
the sample measures the effect of the external magnetic field alone, and the m(H )
dependence is then obtained by integrating and subtracting the signals from both
channels. Dividing Vm (t) by VH (t), one furthermore obtains a signal proportional
to the magnetic susceptibility: VVHm dm
dt : dt = dH ≡ χ .
dH dm

Often, it is possible to move the sample in the gradiometer. Then, in order


to improve signal-to-noise ratio, it is common to measure the background signal
in the very same conditions and use it for the measurement correction. If the
pick-up coil system consists of two identical coils (first-order axial or planar
gradiometer), the sample is consequently measured in either of those, and the
result is averaged. Due to the short pulse duration, temperature control during the
measurement is impossible, so only the field dependencies of the magnetic moment
and susceptibility are determined.
Closed-circuit magnetometry: The most general description of a magnetic system
is obtained by integrating the magnetic flux over the whole space domain. However,
in the presence of a massive body with high relative magnetic permeability, most of
the flux will pass through the body, so the description can be restricted to the body
volume. If this body is shaped as a closed circuit, a major part of electric circuit
formalism, including Kirchhoffs’ laws, can be applied (cf., Sect. “Electromagnets
with Magnetic Core”).
One of the main advantages of using closed circuits for magnetic measurements
is the well-defined flux penetrating the studied specimen, as opposed to open-circuit
methods, where the demagnetizing field can lead to significant distortion of the
internal flux.
The closed-circuit approach is restricted to a moderate field and temperature
range. Yet, it exactly matches the range of interest for characterizing technical
1136 O. Portugall et al.

Fig. 26 Closed sample


circuit

Fig. 27 Sample closed by a


yoke

magnetic materials. Industrial-grade devices are a subject of international standards


(e.g., IEC 60404).
There is a whole variety of commercially available and home-built closed-circuit
devices. One can sort them in two major groups: closed sample devices and sample
closed by yoke devices.
The former class (Fig. 26) is used for testing and characterizing soft magnetic
materials, where the sample itself forms a closed magnetic circuit. The field is
obtained by Ampere’s law H = nI , where I is the current and n is the number of
turns in the primary coil. The full flux density B is measured by the secondary coil.
The outcome of the measurement is a full magnetic hysteresis loop. This type of
device can be operated in both DC and AC mode, depending on the testing protocol.
The second class of devices is based on an electromagnet with a yoke containing
a small air gap (Fig. 27). This gap is partially, or completely, closed by the sample.
The sample shape should match the surface of the yoke poles to prevent spurious
gaps. The flux and the magnetic field H in the sample are measured by dedicated
sensors. The field range in this case is limited by the yoke saturation (including
poles) and typically amounts to about 2–3 MA/m (Fig. 27).

Torque Magnetometers
In an external magnetic field, a magnetic dipole is subject to (i) torque, trying to
orient the dipole along the field direction, and (ii) force, acting in the direction of
the field gradient.

τ = μ0 (m × H)
F = μ0 (m · ∇)H (9)
22 Magnetic Fields and Measurements 1137

Magnetic torque can give a measure of magnetic anisotropy – one of the


key properties of magnetic materials. Among others, it is the origin of magnetic
coercivity and the energy product of permanent magnets. Anisotropy reflects the
tensor nature of magnetic permeability and describes the preference of the sample
magnetization to be oriented along a certain direction or a set of directions. Magnetic
anisotropy can by expressed in terms of anisotropy energy. The specific expression
depends on the crystal symmetry; for instance, for hexagonal crystals, it is given by

EA = K1 cos2 θ + K2 cos4 θ + ..., (10)

where K1 and K2 are anisotropy constants, whose measurement, alongside with


quantifying the magnetic moment, are important tasks of magnetometry.

Magnetic anisometer: The most common way to quantify magnetic anisotropy is to


measure the magnetic moment along different directions. Alternatively, appropriate
devices can measure the magnetic torque, thus accessing the magnetic anisotropy
directly. An example for a torque magnetometer is a setup where the sample is fixed
to a rod placed in a perpendicular magnetic field (Fig. 28). The sample rod transfers
the torque from the sample to a torque sensor. The sample is typically shaped as a
disk, with the plane containing the “hard” magnetization direction. The assembly
can rotate about the rod axis, and the dependence τ (φ) is recorded. Analyzing the
angular dependence, the anisotropy constants K1 and K2 are obtained. The accuracy
of determining K1 is reported to be about 1% [116, 117].
Cantilever magnetometers: If a magnetic sample is fixed to the free end of a
cantilever and exposed to a magnetic field, the cantilever bends according to the
corresponding torque. The torque can originate both from the magnetic torque

Fig. 28 Torque
magnetometer
1138 O. Portugall et al.

experienced by an anisotropic sample in a magnetic field and from a force acting


on a magnetic dipole in a field gradient. Two most common ways of reading out the
bending of a cantilever are capacitive and piezoresistive. Capacitive cantilevers are
made of an elastic metallic foil, which is installed parallel to a conductive substrate
at a distance d so as to form a capacitance C = 0 S/d, where S is the area of
the cantilever pad. If the deflection of the cantilever is small compared to its length,
the movement of the cantilever can be treated as a uniform change of the separation
distance d. The corresponding change of the capacitance C is registered by an LC
bridge [60].
Piezoresistive cantilevers convert the flexing of the cantilever to a relative resis-
tance change of the built-in resistive path. This technique is easy to miniaturize and
suitable for chip production technology. Resistive microcantilevers are produced for
atomic force microscopy and are successfully used as very sensitive magnetometers,
reaching sensitivities of ∼10−13 Am2 [186, 115].
Cantilever magnetometers can top the sensitivity list. A drawback is, however,
their entangled torque and force response. This is why they are primarily used for
detecting quantum oscillations, where this entanglement is not an obstacle. The
separation of torque and force can be done by two successive measurements, one
of which is performed in a strong well-defined field gradient and another one with
the gradient removed [100].
Both capacitive and piezoresistive cantilevers can be equipped with a feedback
circuit, i.e., a single- or multi-turn coil of a defined geometry. The current through
the coil is set such that the cantilever deflection is kept constant. This diminishes
possible nonlinearity effects caused, for instance, by large cantilever deflections.
Cantilevers can reach sensitivities comparable with SQUIDs. Although their
precision can be limited, they can be used in very strong magnetic fields and are
less demanding as far as electric and magnetic shielding is concerned.
A combination of microcantilevers with optical readout and resonance alter-
nating gradient can push the sensitivity even further and yield 10−16 Am2 . A
comprehensive review can be found in [107].
Magnetic scales: One of the most direct methods of measuring force is the balancing
scale, which can be utilized to evaluate magnetic moment by measuring the force
acting on a magnetic sample in a field gradient. This can be implemented in two
types of arrangements, called Faraday and Gouy methods. In the Faraday method,
the sample is placed in a specially formed non-uniform magnetic field with a vertical
field gradient (Fig. 29a). The resulting force then is given by

dH dH
F = μ0 m = μ0 V χ H , (11)
dz dz

where V is the sample volume.


In the Gouy method, only one end of a long sample is placed in the magnetic
field (Fig. 29b). The field in this case is uniform, and the force is the result of the
energy density difference between the magnetized sample and the air gap. If the
sample with cross-sectional area A and susceptibility χ is moved into an air gap
22 Magnetic Fields and Measurements 1139

Fig. 29 Magnetic scales based on the Faraday (a) and Gouy (b) methods

with magnetic field H , the respective change of the magnetic energy U gives rise to
the force

dU 1 d(BH V ) 1 dV 1
F = = = μ0 χ H 2 = μ0 χ H 2 A (12)
dz 2 dz 2 dz 2

The sensitivity of the method is to a large extent defined by the mechanical balance
and can reach ∼10−12 Am2 [121].
Magnetic balance methods have given rise to numerous modifications, one of
the most important being the Evans magnetic balance [47], as its principle is
implemented in a number of commercial devices. There, the force acting on the
sample is counterbalanced in the other arm of the balance by a coil placed in a field
gradient.
Alternating gradient technique: Magnetic scale methods provide good sensitivity,
yet their performance can be spoiled by mechanical noise of the environment. A
typical approach to improve signal-to-noise ratio is modulation of the excitation
parameter and synchronous detection of the signal. Gradient modulation techniques
have first been introduced by Zijlstra [194]. The field gradient is produced by a pair
of coils fed with an AC current, and the force acting on the sample oscillates in
phase with this current. In a vibrating reed magnetometer, the sample is attached to
an elastic reed and forms a cantilever (Fig. 30). The force is read by a strain gauge
attached to the reed. In order to increase the sensitivity, the excitation frequency
is chosen close to the resonance of the system. The sensitivity of conventionally
sized devices (as opposed to microcantilevers from [107]) is about ∼10−9 Am2
[50, 138].

Stray Field Mapping

There are many cases where it is important to map the spatial variation of the stray
field of a magnetic structure. A prominent example is the study of domain walls in
magnetic materials [28]. With the development of magnetic devices for data storage,
1140 O. Portugall et al.

Fig. 30 Vibrating reed magnetometer

actuators, or sensors, the microscopic characterization of magnetic structures has


become even more relevant [132]. For these problems, stray field techniques are
the method of choice. In contrast to bulk magnetic methods for magnetic structure
determination like neutron scattering and magnetic X-ray spectroscopy, stray field
methods explore the conditions at, or close to, the surface. There exist three main
methods for stray field mapping:
Decoration methods use colloidal magnetic particles that accumulate at regions
of maximum magnetic field on a polished surface of a magnetic material. This
technique was developed by Francis Bitter in the 1930s [12]. The underlying
mechanisms are the forces on the magnetic particles generated by the gradient
of the stray magnetic field. The deposited particles can be visualized by a simple
microscope or by more advanced imaging techniques based on magnetic resonance
methods. The spatial resolution of the technique is hence identical with the detection
method (200 nm for microscopy and 10 μm for magnetic resonance methods).
Magnetic force microscopy (MFM) [65] is a common method. Here, the stray field
mapping element is a small ferromagnetic particle located at the tip of a cantilever.
Spatial resolution down to 20 nm is obtained by detecting the force derivative
capacitively, by optical interferometry or the change of the mechanical resonance
frequency, when the cantilever scans the surface of the sample. MFM stray field
techniques can be used down to very low temperatures and may be combined
with other magnetic measurement techniques like inductive or magnetoresistive
detectors and nuclear magnetic resonance. The latter recently led to the development
of magnetic resonance force microscopy (MRFM), a combination of MFM and
magnetic resonance imaging (MRI) [155,195]. This technique allowed the detection
of individual electron spins [145] and the MR imaging of an individual virus with
a resolution of better than 10 nm [34]. When using MFM methods, care has to be
taken that the ferromagnetic tip does not disturb the magnetic structure.
Scanning electron microscopy (SEM) [135, 136] detects magnetic structure either
by measuring the deflection of secondary electrons generated by the stray field
or by detecting their spin polarization. The latter method is known as scanning
22 Magnetic Fields and Measurements 1141

electron microscopy with polarization analysis (SEMPA). Spatial resolution down


to the 10 nm scale is obtained by rastering the beam across the surface. SEM
methods can also be used in combination with decoration techniques. By using a
replica technique, magnetic structures of ferromagnets or the flux line lattice of
superconductors that only occur at low temperatures can be monitored at room
temperature by SEM. The replica of the low temperature structure is obtained with
ferromagnetic particles that are deposited at low temperatures on the surface of the
materials exhibiting a magnetic state. This technique led to the first detection of the
vortex lattice of conventional superconductors [45].
Most recent developments for stray field methods involve SQUIDs [182] and
nitrogen-vacancy (NV) centers in diamonds [148]. The latter technique has the
potential of detecting individual electronic and nuclear magnetic moments.

Calibration and Metrology

The importance of calibrating and re-calibrating experimental equipment is fre-


quently underestimated, and it is a widespread misconception that the quality of
a measurement can be judged from the observed signal fluctuations. As a matter
of fact, the accuracy of most instruments is worse than their precision, and the
difference can easily exceed one order of magnitude. Examples may be found in
the manufacturer’s specifications of NMR probes which are still exemplary insofar
that the distinction between precision and accuracy is actually made. This is not
always the case and correct calibration should therefore be a major concern for any
experiment.
The ultimate accuracy limits of a measurement are determined by primary
standards that are maintained by national or international metrology institutes.
These can be either unique artefacts such as the international prototype kilogram
or normalized measurements of fundamental properties such as the 133 Cs ground-
state hyperfine transition whose radiation frequency determines the unit Second.
Artefacts are chosen by convention, thus making them accurate by definition.
However, their dissemination requires comparisons that are error-prone and hence
limit the transmitted accuracy. Measurement-based standards, on the other hand, are
inherently inaccurate due to their parameter dependence, yet their reproduction is
simpler and more reliable. Their recognition depends on their accepted metrological
quality and is governed by the International Bureau of Standards (BIPM) that
maintains the SI unit system. Parameter values are periodically updated by the
Committee on Data for Science and Technology (CODATA) of the International
Council for Science (ICSU).
On this superior level, the Tesla is determined via the NMR of hydrogen
in water at 298.15 K. Its accuracy is thus limited by the uncertainty of the
proton gyromagnetic ratio γn /(2π ) = 42.57638507(53) MHz/T [106], by error
margins of the frequency measurement, imperfections of the reference material,
and deficiencies in the reproduction of experimental conditions such as temperature
1142 O. Portugall et al.

or the absence of stray fields. The best accuracies that can currently be achieved
under these conditions are of the order of 1 ppm and generally refer to the mT-range
[183,118]. At μT-fields, the relative accuracy inevitably gets worse as the resonance
frequency drops and the screening of stray fields becomes increasingly difficult.
As primary standards exist in highly specialized metrology institutes only, their
dissemination is achieved via different levels of subordinate standards. At the
bottom of the hierarchy are working standards used to calibrate equipment locally,
on a day-to-day basis. These can be either artefacts or measurement devices with a
traceable relationship to the relevant primary standard.
Since NMR has become a relatively conventional and affordable technique, its
use as a working standard is no longer an exception: commercial NMR-based
Teslameters generally promise accuracies of 5–10 ppm. Cheaper and more widely
used are Hall probes which provide less accuracy but have the additional advantage
of small size and a relatively wide field range.
Another alternative that can in principle be homemade are field coils in con-
junction with accurate (and precise) current sources. When designing a field coil
standard, special attention has to be paid to (i) the field homogeneity across the
measurement volume that must be better than the accuracy target; (ii) the choice
of an operating regime that avoids heating and deformation; (iii) the compensation
of parasitic effects associated with, for example, the current leads that connect the
coil; and (iv) the accuracy and precision of the current source. To provide sufficient
homogeneity, most field coil standards are based on long, thin-walled solenoids,
Helmholtz-coils, or more complex Helmholtz-like arrangements such as Braunbek
or Barker coils. For practical reasons, very large specimen may be polygonal rather
than circular [57].
When calibrating one device by another, accuracy is inevitably lost. An interest-
ing study in this respect has been performed by seven European metrology institutes,
each measuring the field-current ratio of the same coil with their respective
instruments and procedures [184]. While the weighted and averaged measurement
uncertainty gave an overall agreement within 50 ppm in this study, individual
results still showed deviations of up to 0.5%. In this respect, it is safe to assume
that a simple, possibly homemade, working standard will usually not permit field
calibrations of much better than 1%. To obtain higher accuracies, only certified
instruments with documented relationship to the respective primary standard should
be used.
Despite their inherent inability to measure absolute field values, SQUIDS
represent a useful tool in metrology. Their unrivaled capacity to detect field vari-
ations with outstanding spatial resolution makes them highly suitable for mapping
inhomogeneities that limit the accuracy of, for example, field coil standards. In
this case, their complete insensibility to a constant background field turns into an
advantage as the offset doesn’t reduce the dynamic range of the measurement.
While Tesla standards are important for calibrating fields and field sensors, they
are of limited use for devices designed to measure absolute magnetic properties. The
preferred technical solution in this case are certified reference materials (CRM).
A CRM for bulk magnetic measurements typically consists of a piece of material
22 Magnetic Fields and Measurements 1143

shaped into a sphere, disc, or cylinder whose magnetic moment or susceptibility


is accurately known. It thus provides a single reference value that can be used to
calibrate the corresponding measurement range of a magnetometer.
In order to provide reliable results, CRMs have to satisfy strict criteria. Their
production and distribution are therefore handled by metrology institutes and
certified industries. A particular concern for end users is questions relating to
stability and aging that determine the storage and expiry of CRMs.
The most widely used CRMs for magnetic measurements are the so-called
Standard Reference Materials (SRM® ) which is a trademark of the National Institute
of Standards and Technology (NIST). The NIST supply includes CRMs for different
measurement ranges and magnetic properties featuring relative uncertainties around
0.3%. Taking into account additional accuracy losses associated with, for example,
calibration errors, this means that, again, normal laboratory experiments will
generally not provide accuracies of better than 1%.

Acknowledgments The authors are grateful to Carlo Rizzo for helpful discussions on metrology
issues.

References
1. A. Abragam. The principles of nuclear magnetism. Clarendon Press, Oxford, 1983.
2. B. Albertazzi, J. Béard, A. Ciardi, T. Vinci, J. Albrecht, J. Billette, T. Burris-Mog, S.N.
Chen, D. Da Silva, S. Dittrich, T. Herrmannsdörfer, B. Hirardin, F. Kroll, M. Nakatsutsumi,
S. Nitsche, C. Riconda, L. Romagnagni, H.-P. Schlenvoigt, S. Simond, E. Veuillot, T.E.
Cowan, O. Portugall, H. Pépin, and J. Fuchs. Production of large volume, strongly magnetized
laser-produced plasmas by use of pulsed external magnetic fields. Rev. Sci. Instrum.,
84:043505, 2013.
3. M.N. Ali, J. Xiong, S. Flynn, J. Tao, Q.D. Gibson, L.M. Schoop, T. Liang, N. Hal-
dolaarachchige, M. Hirschberger, N.P. Ong, and R.J. Cava. Large, non-saturating magne-
toresistance in WTe2. Nature, 514:205–208, 2014.
4. Y. Allain, J. de Gunzbourg, J.P. Krebs, and A. Miedan-Gros. Pulsed field magnetization
measurements up to 500 kOe. Rev. Sci. Instr., 39(9):1360–1365, 1968.
5. G. Aubert. High magnetic fields: physical limits in magnet design. Phys. Scr., T35:168–171,
1991.
6. G. Aubert, L. van Bockstal, E. Fernandez, W. Joss, V. Kuchinski, R. Kurtz, N. Mikhailov,
and Ph. Sala. Quasi-stationary magnetic fields of 60 T using inductive energy storage. IEEE
Trans. Appl. Supercond., 12(1):703–706, 2002.
7. S. Batut, J. Mauchain, R. Battesti, C. Robilliard, M. Fouché, and O. Portugall. A transportable
pulsed magnet system for fundamental investigations in quantum electrodynamics and
particle physics. IEEE Trans. Appl. Supercond., 18(2):600–603, 2008.
8. M.D. Bird. Resistive magnet technology for hybrid inserts. Supercond. Sci. Technol.,
17(8):R19–R33, 2004.
9. M.D. Bird, S. Bole, Y.M. Eyssa, B.J. Gao, and H.J. Schneider-Muntau. Design of a poly-bitter
magnet at the NHMFL. IEEE Trans. Magn., 32(4):2542–2545, 1996.
10. M.D. Bird, I.R. Dixon, and J. Toth. Design of the next generation of florida-bitter magnets at
the NHMFL. IEEE Trans. Appl. Supercond., 14(2):1253–1256, 2004.
11. M.D. Bird, I.R. Dixon, and J. Toth. Large, high-field magnet projects at the NHMFL. IEEE
Trans. Appl. Supercond., 25(3):1–6, 2015.
1144 O. Portugall et al.

12. F. Bitter. On inhomogeneities in the magnetization of ferromagnetic materials. Phys. Rev.,


38(10):1903–1905, 1931.
13. F. Bitter. The design of powerful electromagnets part I. the use of iron. Rev. Sci. Instr.,
7(12):479–481, 1936.
14. F. Bitter. The design of powerful electromagnets part III. the use of iron. Rev. Sci. Instr.,
8(9):318–319, 1937.
15. F. Bitter. Water cooled magnets. Rev. Sci. Instrum., 33(3):342, 1962.
16. R.C. Black, F.C. Wellstood, E. Dantsker, A.H. Miklich, D. Koelle, F. Ludwig, and J. Clarke.
High-frequency magnetic microscopy using a high-Tc SQUID. IEEE Trans. Appl. Super-
cond., 5(2):2137–2141, 1995.
17. L. Van Bockstal, G. Heremans, and F. Herlach. Coils with fibre composite reinforcement for
pulsed magnetic fields in the 50-70 T range. Meas. Sci. Technol., 2:1159–1164, 1991.
18. G.J. Bowden. Detection coil systems for vibrating sample magnetometers. J. Phys. E: Sci.
Instrum., 5(11):1115, 1972.
19. B.A. Boyko, A.I. Bykov, M.I. Dolotenko, N.P. Kolokol’chikov, I.M. Markevtsev, O.M.
Tatsenko, and A.M. Shuvalov. More than 20 MG magnetic field generation in the cascade
magnetocumulative MC-1 generator. In H.J. Schneider-Muntau, editor, Proc. 8th Int. Conf.
Megagauss Magnetic Field Generation and Related Topics, pages 61–66. World Scientific,
Singapore, 1999.
20. J. Chavanne and G. Le Bec. Prospects for the use of Permanent Magnets in Future Accelerator
Facilities. In Proceedings, 5th International Particle Accelerator Conference (IPAC 2014),
page TUZB01, 2014.
21. W.G. Chen, Y.F. Tan, Z.M. Chen, J.W. Zhu, Z.Y. Chen, Y.N. Pan, F.T. Wang, P.C. Huang, and
G.L. Kuang. Final design of the 40 T hybrid magnet superconducting outsert. IEEE Trans.
Appl. Supercond., 23(3):4300404–4300404, 2013.
22. D. Ciudad, C. Aroca, M.C. Sánchez, E. Lopez, and P. Sánchez. Modeling and fabrication of
a MEMS magnetostatic magnetic sensor. Sens. Actuators A Phys., 115:408–416, 2004.
23. W.G. Clark, T.W. Hijmans, and W.H. Wong. Multiple coil pulsed NMR method for measuring
the multipole moments of particle accelerator bending magnets. J. Appl. Phys., 63(8):4185,
1988.
24. J. Clarke. Squids: Principles, noise, and applications. In S.T. Ruggiero and D.A. Rudman,
editors, Superconducting Devices, pages 51 – 99. Academic Press, 1990.
25. J. Clarke and A.I. Braginski (Eds.). The SQUID handbook, volume 1. Wiley-Vch, 2004.
26. E.C. Cnare. Magnetic flux compression by magnetically imploded metallic foils. J. Appl.
Phys., 37(10):3812, 1966.
27. J.M.D. Coey. Permanent magnet applications. J. Magn. Magn. Mater., 248(3):441–456,
2002.
28. J.M.D. Coey. Magnetism and Magnetic Materials, page 355. Cambridge University Press
(CUP), 2009.
29. C. Coillot and P. Leroy. Induction magnetometers principle, modeling and ways of
improvement. In Magnetic Sensors - Principles and Applications. InTech, 2012.
30. F.C.S. da Silva, W.C. Uhlig, A.B. Kos, S. Schima, J. Aumentado, J. Unguris, and D.P. Pappas.
Zigzag-shaped magnetic sensors. Appl. Phys. Lett., 85(24):6022–6024, 2004.
31. E. Danieli, J. Mauler, J. Perlo, B. Blümich, and F. Casanova. Mobile sensor for high resolution
NMR spectroscopy and imaging. J. Magn. Reson., 198(1):80–87, 2009.
32. B.S. Deaver and W.M. Fairbank. Experimental Evidence for Quantized Flux in Supercon-
ducting Cyclinders. Phys. Rev. Lett., 7:43–46, 1961.
33. F. Debray and P. Frings. State of the art and developments of high field magnets at the
“laboratoire national des champs magnétiques intenses”. C. R. Phys., 14(1):2–14, 2013.
34. C.L. Degen, M. Poggio, H.J. Mamin, C.T. Rettner, and D. Rugar. Nanoscale magnetic
resonance imaging. Proc. Natl. Acad. Sci. U.S.A., 106(5):1313–1317, 2009.
35. D. DiLella, L.J. Whitman, R.J. Colton, T.W. Kenny, W.J. Kaiser, E.C. Vote, J.A. Podosek,
and L.M. Miller. A micromachined magnetic-field sensor based on an electron tunneling
displacement transducer. Sens. Actuators A Phys., 86:8–20, 2000.
22 Magnetic Fields and Measurements 1145

36. I.R. Dixon, T.A. Adkins, M.D. Bird, S.T. Bole, J.Toth, H.Ehmler, M.Hoffman, and P.Smeibidl.
Final assembly of the Helmholtz-Zentrum Berlin series-connected hybrid magnet system.
IEEE Trans. Appl. Supercond., 25(3):1–4, 2015.
37. E. Dormann, G. Sachs, W. Stöcklein, B. Bail, and M. Schwoerer. Gaussmeter application of
an organic conductor. Appl. Phys. A, 30(4):227–231, 1983.
38. V. Druzhinin, O.M. Tatsenko, A.I. Bykov, M.I. Dolotenko, N.P. Kolokol’chikov, Y.B.
Kudasov, V.M. Platonov, C.M. Fowler, B.L. Freeman, J.D. Goettee, J.C. King, W. Lewis,
B.R. Marshall, B.J. Papatheofanis, P.J. Rodriguez, L.R. Veeser, and W.D. Zerwekh. Nonlinear
Faraday effect in CdS semiconductor in an ultrahigh magnetic field. Physica B, 211:392–395,
1995.
39. E. du Trémolet de Lacheisserie, D. Gignoux, and M. Schlenker. Permanent magnets. In
Magnetism, pages 3–88. Springer Science + Business Media, 2002.
40. F. Duc, X. Fabrèges, T. Roth, C. Detlefs, P. Frings, M. Nardone, J. Billette, M. Lesourd,
L. Zhang, A. Zitouni, P. Delescluse, J. Béard, J. P. Nicolin, and G.L.J.A. Rikken. A 31 T
split-pair pulsed magnet for single crystal x-ray diffraction at low temperature. Rev. Sci.
Instrum., 85:053905, 2014.
41. D. Dufeu, E. Eyraud, and P. Lethuillier. An efficient 8 T extraction vector magnetometer with
sample rotation for routine operation. Rev. Sci. Instrum., 71(2):458, 2000.
42. H.B. Dwight and C.F. Abt. The shape of core for laboratory electromagnets. Rev. Sci. Instr.,
7(3):144–146, 1936.
43. A.S. Edelstein. Advances in magnetometry. J. Phys. Condens. Matter, 19:165217, 2007.
44. A.S. Edelstein, G.A. Fischer, M. Pedersen, E.R. Nowak, Shu Fan Cheng, and C.A. Nordman.
Progress toward a thousandfold reduction in 1/f noise in magnetic sensors using an ac
microelectromechanical system flux concentrator. J. Appl. Phys., 99:08B317, 2006.
45. U. Essmann and H. Träuble. The direct observation of individual flux lines in type II
superconductors. Phys. Lett. A, 24(10):526–527, 1967.
46. K.P. Estola and J. Malmivuo. Air-core induction-coil magnetometer design. J. Phys. E: Sci.
Instrum., 15:1110–1113, 1982.
47. D.F. Evans. A new type of magnetic balance. J. Phys. E: Sci. Instrum., 7(4):247, 1974.
48. J.A. Ewing and W. Low. On the magnetisation of iron and other magnetic metals in very
strong fields. Philos. Trans. A: Math. Phys. Eng. Sci., 180(0):221–244, 1889.
49. P. Fazilleau, C. Berriaud, R. Berthier, F. Debray, B. Hervieu, W. Joss, F.P. Juster,
M. Massinger, C. Mayri, Y. Queinec, C. Pes, R. Pfister, P. Pugnat, L. Ronayette, and
C. Trophime. Final design of the new Grenoble hybrid magnet. IEEE Trans. Appl. Supercond.,
22(3):4300904–4300904, 2012.
50. P.J. Flanders. A vertical force alternating-gradient magnetometer. Rev. Sci. Instr., 61(2):839–
847, 1990.
51. S. Foner. Versatile and sensitive vibrating-sample magnetometer. Rev. Sci. Instrum.,
30(7):548, 1959.
52. P. Frings, J. Vanacken, C. Detlefs, F. Duc, J.E. Lorenzo, M. Nardone, J. Billette, A. Zitouni,
W. Bras, and G.L.J.A. Rikken. Synchrotron x-ray powder diffraction studies in pulsed
magnetic fields. Rev. Sci. Instrum., 77:063903, 2006.
53. P. Frings, H. Witte, H. Jones, J. Béard, and Thomas Hermannsdörfer. Rapid cooling methods
for pulsed magnets. IEEE Trans. Appl. Supercond., 18(2):612–615, 2008.
54. E. Fukushima and S. Roeder. Experimental pulse NMR : a nuts and bolts approach. Addison-
Wesley Pub. Co., Advanced Book Program, Reading, Mass, 1981.
55. Zhehong Gan, I. Hung, Xiaoling Wang, J. Paulino, Gang Wu, I.M. Litvak, P.L. Gor'kov, W.W.
Brey, P. Lendi, J.L. Schiano, M.D. Bird, I.R. Dixon, J. Toth, G.S. Boebinger, and T.A. Cross.
NMR spectroscopy up to 35.2 T using a series-connected hybrid magnet. J. Magn. Reson.,
284:125–136, 2017.
56. M.W. Garrett. Calculation of fields, forces, and mutual inductances of current systems by
elliptic integrals. J. Appl. Phys., 34(9):2567–2573, 1963.
57. M.W. Garrett and S. Pissanetzky. Polygonal coil systems for magnetic fields with homogene-
ity of the fourth to the eighth order. Rev. Sci. Instr., 42(6):840–857, 1971.
1146 O. Portugall et al.

58. H. H. Gatzen, E. Andreeva, and H. Iswahjudi. Eddy-current microsensor based on thin-film


technology. IEEE Trans. Magn., 38:3368–3370, 2002.
59. C. Gemmel, W. Heil, S. Karpuk, K. Lenz, Ch. Ludwig, Yu. Sobolev, K. Tullney, M. Burghoff,
W. Kilian, S. Knappe-Grüneberg, W. Müller, A. Schnabel, F. Seifert, L. Trahms, and St.
Baeßler. Ultra-sensitive magnetometry based on free precession of nuclear spins. Eur. Phys.
J. D, 57(3):303–320, 2010.
60. R. Griessen. A capacitance torquemeter for de Haas-van Alphen measurements. Cryogenics,
13(6):375–377, 1973.
61. M. Grisi, G.M. Conley, P. Sommer, J. Tinembart, and G. Boero. A single-chip integrated
transceiver for high field NMR magnetometry. Rev. Sci. Instrum., 90(1):015001, 2019.
62. Seungyong Hahn, Kwanglok Kim, Kwangmin Kim, Xinbo Hu, T Painter, I Dixon, Seokho
Kim, KR. Bhattarai, S Noguchi, J Jaroszynski, and D.C. Larbalestier. 45.5-Tesla direct-
current magnetic field generated with a high-temperature superconducting magnet. Nature,
570(7762):496–499, 2019.
63. Hyung-Suk Han and Dong-Sung Kim. Magnetic Levitation. Springer Netherlands, 2016.
64. R.K. Harris, E.D. Becker, S.M. Cabral de Menezes, P.Granger, R.E. Hoffman, and K.W. Zilm.
Further conventions for NMR shielding and chemical shifts (IUPAC recommendations 2008).
Pure Appl. Chem., 80(1), 2008.
65. U. Hartmann. Magnetic force microscopy. Annu. Rev. Mater. Sci., 29(1):53–87, 1999.
66. K. Hashi, S. Ohki, S. Matsumoto, G. Nishijima, A. Goto, K. Deguchi, K. Yamada, T. Noguchi,
S. Sakai, M. Takahashi, Y. Yanagisawa, S. Iguchi, T. Yamazaki, H. Maeda, R. Tanaka,
T. Nemoto, H. Suematsu, T. Miki, K. Saito, and T. Shimizu. Achievement of 1020 MHz
NMR. J. Magn. Reson., 256:30–33, 2015.
67. F. Herlach. Pulsed magnets. Rep. Prog. Phys., 62:859–920, 1999.
68. F. Herlach and R. McBroom. Megagauss fields in single turn coils. J. Phys. E: Sci. Instrum.,
6:652–654, 1973.
69. A.L. Herrera-May, L.A. Aguilera-Cortés, P.J. García-Ramírez, and E. Manjarrez. Resonant
magnetic field sensors based on MEMS technology. Sensors, 9:7785–7713, 2009.
70. High Field Magnet Laboratory (HFML), Radboud University (RU), Nijmegen, The Nether-
lands. http://www.ru.nl/hfml.
71. High Magnetic Field laboratory (HMFL), Chinese Academy of Science (CAS), Hefei, China.
http://english.hf.cas.cn/r/ResearchDivisions/HFML.
72. D.R. Hines, S.A. Solin, T. Thio, and T. Zhou. Extraordinary magnetoresistance at room
temperature in inhomogeneous narrow-gap semiconductors, 2004. US Patent 6,707,122.
73. Hochfeld-Magnetlabor (HLD), Helmholtz Zentrum Dresden Rossendorf (HZDR), Dresden,
Germany. http://www.hzdr.de/hld.
74. Y. Honkura. Development of amorphous wire type MI sensors for automobile use. J. Magn.
Magn. Mater., 249(1-2):375–381, 2002.
75. S.R. Hoon and S.N.M. Willcock. The design and operation of an automated double-crank
vibrating sample magnetometer. J. Phys. E: Sci. Instrum., 21(8):772, 1988.
76. D.I. Hoult and R.E. Richards. The signal-to-noise ratio of the nuclear magnetic resonance
experiment. J. Magn. Reson., 213(2):329–343, 2011.
77. D.I Hoult, R.E Richards, and P Styles. A novel field-frequency lock for a superconducting
spectrometer. J. Magn. Reson., 30(2):351–365, 1978.
78. N. Ida. Engineering Electromagnetics. Springer New York, New York, 2000.
79. G.T. Noe II, H. Nojiri, J. Lee, G.L. Woods, J. Léotin, and J. Kono. A table-top, repetitive
pulsed magnet for nonlinear and ultrafast spectroscopy in high magnetic fields up to 30 T.
Rev. Sci. Instrum., 84:123906, 2013.
80. International Megagauss Science Laboratory (IMGSL), Institute for Solid State Physics
(ISSP), Chiba, Japan. http://issp.u-tokyo.ac.jp/labs/mgsl.
81. Z. Islam, J.P.C. Ruff, H. Nojiri, Y.H. Matsuda, K.A. Ross, B.D. Gaulin, Zhe Qu, , and J.C.
Lang. A portable high-field pulsed-magnet system for single-crystal x-ray scattering studies.
Rev. Sci. Instrum., 80:113902, 2009.
82. J.H. Iwamiya and S.W. Sinton. Stray-field magnetic resonance imaging of solid materials.
Solid State Nucl. Magn. Reson., 6(4):333–345, 1996.
22 Magnetic Fields and Measurements 1147

83. R. Jahns, R. Knöchel, H. Greve, E. Woltermann, E. Lage, and E. Quandt. Magnetoelectric


sensors for biomagnetic measurements. In 2011 IEEE International Symposium on Medical
Measurements and Applications, page 107. Institute of Electrical and Electronics Engineers
(IEEE), 2011.
84. J.T. Jeng, C.Y. Chiang, C.H. Chang, and C.C. Lu. Vector magnetometer with dual-bridge
GMR sensors. IEEE Trans. Magn., 50(1):1–4, 2014.
85. F. Jiang, T. Peng, H. Xiao, J. Zhao, Y. Pan, F. Herlach, and L. Li. Design and test of a flat-top
magnetic field system driven by capacitor banks. Rev. Sci. Instrum., 85:045106, 2014.
86. L. Jogschies, D. Klaas, R. Kruppe, J. Rittinger, P. Taptimthong, A. Wienecke, L. Rissing,
and M. Wurz. Recent developments of magnetoresistive sensors for industrial applications.
Sensors, 15(11):28665–28689, 2015.
87. B.D. Josephson. The discovery of tunnelling supercurrents. Rev. mod. Phys., 46:251–254,
1974.
88. F. Keplinger, S. Kvasnica, A. Jachimowicz, F. Kohl, J. Steurer, and H. Hauser. Lorentz force
based magnetic field sensor with optical readout. Sens. Actuators A Phys., 110:112–118,
2004.
89. K. Kindo. New pulsed-magnets for 100 T, long-pulse and diffraction measurements. J. Phys.
Conf. Ser., 51:118, 2006.
90. K. Kindo, S. Takeyama, M. Tokunaga, Y.H. Matsuda, E. Kojima, A. Matsuo, K. Kawaguchi,
and H. Sawabe. The International MegaGauss Laboratory at ISSP, The University of Tokyo.
J. Low. Temp. Phys., 159:381–388, 2010.
91. K.H. Knoepfel. Magnetic Fields. John Wiley & Sons, New York, 2000.
92. Laboratoire National des Champs Magnétiques Intenses (LNCMI), Centre National
de la Recherche Scientifique (CNRS), Toulouse & Grenoble, France. http://www.lncmi.cnrs.
fr.
93. P.J. Lee. Engineering critical current density vs. applied field for superconductors available
in long lengths. https://nationalmaglab.org/magnet-development/applied-superconductivity-
center/plots, 2018. This link contains actual values of critical currents of superconductors and
their bibliographic references.
94. J. Lenz and A.S. Edelstein. Magnetic sensors and their applications. IEEE Sens. J., 6(3):631–
649, 2006.
95. S.H. Liou, X. Yin, S.E. Russek, R. Heindl, F.C.S. Da Silva, J. Moreland, D.P. Pappas, L. Yuan,
and J. Shen. Picotesla magnetic sensors for low-frequency applications. IEEE Trans. Magn.,
47(10):3740–3743, 2011.
96. K. Mackay, M. Bonfim, D. Givord, and A. Fontaine. 50 T pulsed magnetic fields in
microcoils. J. Appl. Phys., 87:1996, 2000.
97. D.J. Mapps. Magnetoresistive sensors. Sens. Actuators A Phys., 59(1-3):9–19, 1997.
98. S. Marauska, R. Jahns, H. Greve, E. Quandt, R. Knöchel, and B. Wagner. MEMS magnetic
field sensor based on magnetoelectric composites. J. Micromech. Microeng., 22:065024–
065030, 2012.
99. Y.H. Matsuda, F. Herlach, S. Ikeda, and N. Miura. Generation of 600 T by electromagnetic
flux compression with improved implosion symmetry. Rev. Sci. Instrum., 73:4288–4294,
2002.
100. A. McCollam, P.G. van Rhee, J. Rook, E. Kampert, U. Zeitler, and J.C. Maan. High sensitivity
magnetometer for measuring the isotropic and anisotropic magnetisation of small samples.
Rev. Sci. Instr., 82(5):053909, 2011.
101. T. McGuire and R. Potter. Anisotropic magnetoresistance in ferromagnetic 3d alloys. IEEE
Trans. Magn., 11(4):1018–1038, 1975.
102. B. Meier, S. Greiser, J. Haase, T. Herrmannsdörfer, F. Wolff-Fabris, and J. Wosnitza. NMR
signal averaging in 62 T pulsed fields. J. Magn. Reson., 210(1):1–6, 2011.
103. C.H. Mielke and R.D. McDonald. Single turn multi-megagauss system at the nhmfl-los
alamos to study plutonium. In 2006 IEEE Int. Conf. on Megagauss Magnetic Field Generation
and Related Topics, pages 227–231. IEEE, 2008.
104. J.R. Miller. The NHMFL 45-T hybrid magnet system: past, present, and future. IEEE Trans.
Appl. Supercond., 13(2):1385–1390, 2003.
1148 O. Portugall et al.

105. N. Mitchell, D. Bessette, R. Gallix, C. Jong, J. Knaster, P. Libeyre, C. Sborchia, and F. Simon.
The ITER magnet system. IEEE Trans. Appl. Supercond., 18(2):435–440, 2008.
106. P.J. Mohr, D.B. Newell, and B.N. Taylor. CODATA recommended values of the fundamental
physical constants: 2014. J. Phys. Chem. Ref. Data, 45(4):043102, 2016.
107. J. Moreland. Micromechanical instruments for ferromagnetic measurements. J. Phys. D:
Appl. Phys., 36(5):R39, 2003.
108. D. Nakamura, A. Ikeda, H. Sawabe, Y. H. Matsuda, and S. Takeyama. Record indoor magnetic
field of 1200 T generated by electromagnetic flux-compression. Rev. Sci. Instr., 89(9):095106,
2018.
109. K. Nakao, F. Herlach, T. Goto, S. Takeyama, T. Sakakibara, and N. Miura. A laboratory
instrument for generating magnetic fields over 200 T with single turn coils. J. Phys. E: Sci.
Instrum., 18:1018, 1985.
110. National High Magnetic Field Laboratory (NHMFL), Tallahassee, Gainesville & Los Alamos,
USA. https://nationalmaglab.org.
111. National High Magnetic Field Laboratory (NHMFL). 32 Tesla All-Superconducting Mag-
net. https://nationalmaglab.org/magnet-development/magnet-science-technology/magnet-
projects/32-tesla-scm, 2017.
112. A. Nikiel, P. Blümler, W. Heil, M. Hehn, S. Karpuk, A. Maul, E. Otten, L.M. Schreiber, and
M. Terekhov. Ultrasensitive 3He magnetometer for measurements of high magnetic fields.
Eur. Phys. J. D, 68(11), 2014.
113. S. Nimori. Present status of the Tsukuba magnet laboratory. J. Low Temp. Phys., 177(1-
2):80–89, 2012.
114. S. Nimori and G. Kido. Current status of the Tsukuba magnet laboratory. J. Low Temp. Phys.,
159(1-2):358–365, 2010.
115. E. Ohmichi and T. Osada. Torque magnetometry in pulsed magnetic fields with use of a
commercial microcantilever. Rev. Sci. Instr., 73(8):3022, 2002.
116. F. Ono, Y. Ohtsu, and O. Yamada. Determination of magnetic anisotropy constants from
unsaturated torque curves in Nd2Fe14B. J. Phys. Soc. Jpn., 55(11):4014–4019, 1986.
117. Y. Otani, H. Miyajima, and S. Chikazumi. High field torque magnetometer for superconduct-
ing magnets. Jpn. J. Appl. Phys., 26(4R):623, 1987.
118. P. Park, Y. Kim, and V. Shifrin. Maintenance of magnetic flux density standards on the basis
of proton gyromagnetic ratio at KRJSS. In 2004 Conference on Precision Electromagnetic
Measurements, page 504. Institute of Electrical and Electronics Engineers (IEEE), 2004.
119. M.A. Paun, J.M. Sallese, and M. Kayal. Hall effect sensors design, integration and behavior
analysis. J. Sens. Actuator Netw., 2(1):85–97, 2013.
120. J. Perlo, F. Casanova, and B. Blümich. Ex situ NMR in highly homogeneous fields: 1H
spectroscopy. Science, 315(5815):1110–1112, 2007.
121. L. Petersson and A. Ehrenberg. Highly sensitive Faraday balance for magnetic susceptibility
studies of dilute protein solutions. Rev. Sci. Instr., 56(4):575–580, 1985.
122. M.H. Phan and H.X. Peng. Giant magnetoimpedance materials: Fundamentals and applica-
tions. Prog. Mater. Sci., 53(2):323–420, 2008.
123. T. Polenova and T.F. Budinger. Ultrahigh field NMR and MRI: Science at a crossroads.
report on a jointly-funded NSF, NIH and DOE workshop, held on november 12–13, 2015 in
bethesda, maryland, USA. J. Magn. Reson., 266:81–86, 2016.
124. R.S. Popovic. High resolution Hall magnetic sensors. In 2014 29th International Conference
on Microelectronics Proceedings - MIEL 2014, pages 69–74. Institute of Electrical &
Electronics Engineers (IEEE), 2014.
125. O. Portugall, N. Puhlmann, H.U. Müller, M. Barczewski, I. Stolpe, and M. von Ortenberg.
Megagauss magnetic field generation in single-turn coils: new frontiers for scientific experi-
ments. J. Phys. D: Appl. Phys., 32:2354–2366, 1999.
126. H. Primas and H.H. Günthard. Field stabilizer for high-resolution nuclear magnetic
resonance. Rev. Sci. Instrum., 28(7):510, 1957.
127. F. Primdahl. The fluxgate magnetometer. J. Phys. E: Sci. Instrum., 12:241–253, 1979.
128. P. Pugnat and H. J. Schneider-Muntau. Conceptual design optimization of a 60 T hybrid
magnet. IEEE Trans. Appl. Supercond., 30(4):1–7, 2020.
22 Magnetic Fields and Measurements 1149

129. P. Pugnat and H.J. Schneider-Muntau. Hybrid magnets: Past, present, and future. IEEE Trans.
Appl. Supercond., 24(3):1–6, 2014.
130. N. Puhlmann, O. Portugall, H.U. Müller, M. Barczewski, I. Stolpe, and M. von Ortenberg.
Solid state applications of a transportable low-cost megagauss generator. J. Phys. D: Appl.
Phys., 30:1861–1866, 1997.
131. Quantum Design, Inc., 6325 Lusk Boulevard, San Diego, USA. PPMS, ACMS option.
132. M. Rahm, J. Raabe, R. Pulwey, J. Biberger, W. Wegscheider, D. Weiss, and C. Meier. Planar
Hall sensors for micro-Hall magnetometry. J. Appl. Phys., 91(10):7980, 2002.
133. H. Raich and P. Blümler. Design and construction of a dipolar halbach array with a
homogeneous field from identical bar magnets: NMR mandhalas. Concepts Magn. Reson.
Part B: Magn. Reson. Eng., 23B(1):16–25, 2004.
134. A.P. Ramirez. Colossal magnetoresistance. J. Phys.: Condens. Matter, 9(39):8171–8199,
1997.
135. E.I. Rau and G.V. Spivak. Scanning electron microscopy of two-dimensional magnetic stray
fields. Scanning, 3(1):27–34, 1980.
136. L. Reimer. Scanning Electron Microscopy. Springer Science + Business Media, 1985.
137. C. Reymond. Magnetic resonance techniques. http://doi.org/10.5170/CERN-1998-005.219,
1998.
138. H.J. Richter, K.A. Hempel, and J. Pfeiffer. Improvement of sensitivity of the vibrating reed
magnetometer. Rev. Sci. Instr., 59(8):1388–1393, 1988.
139. P. Ripka. Review of fluxgate sensors. Sens. Actuators A Phys., 33(3):129–141, 1992.
140. P. Ripka. Induction sensors. In P. Ripka, editor, Magnetic sensors and magnetometers,
chapter 2, pages 47–73. Artech House, 2001.
141. P. Ripka, editor. Magnetic sensors and magnetometers. Artech House, Norwood, MA, 2001.
142. P. Ripka. Advances in fluxgate sensors. Sens. Actuators A Phys., 106(1-3):8–14, 2003.
143. P. Ripka and M. Janošek. Advances in magnetic field sensors. IEEE Sens. J., 10(6):1108–
1116, 2010.
144. L.G. Rubin, B.L. Brandt, R.J. Weggel, S. Foner, and E.J. McNiff. 33.6 T DC magnetic field
produced in a hybrid magnet with Ho pole pieces. Appl. Phys. Lett., 49(1):49–51, 1986.
145. D. Rugar, R. Budakian, H.J. Mamin, and B.W. Chui. Single spin detection by magnetic
resonance force microscopy. Nature, 430(6997):329–332, 2004.
146. A.A. Samoilenko, D.Yu Artemow, and L.A. Sibel’dina. Formation of sensitive layer in
experiments on NMR subsurface imaging of solids. JETP Letters, 47:417, 1988.
147. J.B. Schillig, H.J. Boenig, J.D. Rogers, and J.R. Sims. Design of a 400 MW power supply for
a 60 T pulsed magnet. IEEE Trans. Magn., 30(4):1170–1173, 1994.
148. R. Schirhagl, K. Chang, M. Loretz, and C.L. Degen. Nitrogen-vacancy centers in diamond:
Nanoscale sensors for physics and biology. Annu. Rev. Phys. Chem., 65(1):83–105, 2014.
149. H.J. Schneider-Muntau. Polyhelix magnets. IEEE Trans. Magn., 17(5):1775–1778, 1981.
150. H.J. Schneider-Muntau, J.C. Picoche, P. Rub, and J.C. Vallier. The generation of high
magnetic fields in Grenoble — installations and magnets. In Application of High Magnetic
Fields in Semiconductor Physics, pages 531–541. Springer Science + Business Media,
1982.
151. H.J. Schneider-Muntau and S. Prestemon. Poly-bitter magnets. IEEE Trans. Magn.,
28(1):486–488, 1992.
152. H Schwalbe. Editorial: New 1.2 GHz NMR spectrometers- new horizons? Angewandte
Chemie International Edition, 56(35):10252–10253, 2017.
153. B. Seeber. Handbook of applied superconductivity, chapter 2, page 1195. Institute of Physics
Pub, Bristol Philadelphia, 1998.
154. J.W. Shearer. Interaction of capacitor-bank-produced megagauss magnetic field with small
single-turn coil. J. Appl. Phys., 40(11):4490–4497, 1969.
155. J.A. Sidles. Noninductive detection of single-proton magnetic resonance. Appl. Phys. Lett.,
58(24):2854, 1991.
156. R. Skomski. Permanent magnets: History, current research, and outlook. In Novel Functional
Magnetic Materials, pages 359–395. Springer Science + Business Media, 2016.
157. C. Slichter. Principles of magnetic resonance. Springer-Verlag, Berlin, New York, 1990.
1150 O. Portugall et al.

158. D.O. Smith. Development of a vibrating-coil magnetometer. Rev. Sci. Instrum., 27(5):261,
1956.
159. S.A. Solin, T. Thio, D. R. Hines, and J.J. Heremans. Enhanced room-temperature geometric
magnetoresistance in inhomogeneous narrow-gap semiconductors. Science, 289(5484):1530–
1532, 2000.
160. Jung-Bin Song, X. Chaud, B. Borgnic, F. Debray, P. Fazilleau, and T. Lecrevisse. Construction
and test of a 7 T metal-as-insulation HTS insert under a 20 T high background magnetic field
at 4.2 K. IEEE Trans. Appl. Supercond., 29(5):1–5, 2019.
161. Jung-Bin Song, X. Chaud, F. Debray, P. Fazilleau, and T. Lécrevisse. The high field HTS
insert NOUGAT reached a record field of 32.5 T. https://emfl.eu/the-high-field-hts-insert-
nougat-reached-a-record-field-of-32-5-t/, 2019.
162. K. Spencer, F. Lecouturier, L. Thilly, and J.D. Embury. Established and emerging materials
for use as high-field magnet conductors. Adv. Eng. Mater., 6(5):290–297, 2004.
163. J. Stratton. Electromagnetic theory. McGraw-Hill Book Company, New york, 1941.
164. W. F. Stuart. Earth’s field magnetometry. Rep. Prog. Phys., 35:803–881, 1972.
165. N.A. Stutzke, S.E. Russek, D.P. Pappas, and M. Tondra. Low-frequency noise measurements
on commercial magnetoresistive magnetic field sensors. J. Appl. Phys., 97(10):10Q107, 2005.
166. L. Sun, S. Jiang, and J.R. Marciante. All-fiber optical magnetic-field sensor based on Faraday
rotation in highly terbium-doped fiber. Optics Express, 18(5):5407–5412, 2010.
167. S.M. Thompson. The discovery, development and future of GMR: The Nobel Prize 2007. J.
Phys. D: Appl. Phys., 41(9):093001, 2008.
168. M. Tinkham. Introduction to Superconductivity: Second Edition. Dover Books on Physics.
Dover Publications, 2004.
169. M.T. Todaro, L. Sileo, and M. De Vittorio. Magnetic field sensors based on microelectrome-
chanical systems (MEMS) technology. In Magnetic Sensors - Principles and Applications.
InTech, 2012.
170. Tristan Technologies Inc. http://tristantech.com/sensors.
171. C. Trophime, K. Egorov, F. Debray, W. Joss, and G. Aubert. Magnet calculations at the
Grenoble High Magnetic Field Laboratory. IEEE Trans. Appl. Supercond., 12(1):1483–1487,
2002.
172. C. Trophime, S. Kramer, and G. Aubert. Magnetic field homogeneity optimization of the
giga-NMR resistive insert. IEEE Trans. Appl. Supercond., 16(2):1509–1512, 2006.
173. Tsukuba Magnet Laboratory (TML), National Institute for Materials Science (NIMS),
Tsukuba, Japan. http://www.nims.go.jp/TML/english/.
174. S. Tumanski. Induction coil sensors - a review. Meas. Sci. Technol., 18:R31–R46, 2007.
175. S. Tumanski. Handbook of magnetic measurements. CRC Press, Boca raton, FL, 2011.
176. S. Tumanski. Modern magnetic field sensors — a review. Przeglad Elektrotechniczny, ISSN
0033-2097, 10:R 89, 2013. (http://pe.org.pl/articles/2013/10/1.pdf).
177. A. Balbin Villaverde, D.A. Donatti, and D.G. Bozinis. Terbium gallium garnet verdet constant
measurements with pulsed magnetic field. J. Phys. C: Solid State Phys., 11:L495–L498, 1978.
178. M. von Ortenberg, N. Puhlmann, I. Stolpe, H.U. Müller, A. Kirste, and O. Portugall. The
Humboldt High Magnetic Field Center at Berlin. Physica B, 294-295:568–573, 2001.
179. Y. Wang, J. Li, and D. Viehland. Magnetoelectrics for magnetic sensor applications: status,
challenges and perspectives. Mater. Today, 17(6):269–275, 2014.
180. H. Weber, H.J. Schneider-Muntau, and G. Landwehr. Optimization calculations of polyhelix
coils for hybrid magnets. Appl. Phys., 20(2):163–169, 1979.
181. H.W. Weijers, W.D. Markiewicz, A.J. Voran, S.R. Gundlach, W.R. Sheppard, B. Jarvis,
Z.L. Johnson, P.D. Noyes, J. Lu, H. Kandel, H. Bai, A.V. Gavrilin, Y.L. Viouchkov, D.C.
Larbalestier, and D.V. Abraimov. Progress in the development of a superconducting 32 T
magnet with REBCO high field coils. IEEE Trans. Appl. Supercond., 24(3):1–5, 2014.
22 Magnetic Fields and Measurements 1151

182. W. Wernsdorfer. From micro- to nano-SQUIDs: applications to nanomagnetism. Supercond.


Sci. Technol., 22(6):064013, 2009.
183. K. Weyand. Maintenance and dissemination of the magnetic field unit at PTB. IEEE Trans.
Instrum. Meas., 50(2):470–473, 2001.
184. K. Weyand, E. Simon, L. Henderson, J. Bartholomew, G.M. Teunisse, I. Blanc, G. Crotti,
J. Kupec, A. Jeglic, G. Gersak, J. Humar, and L. Puranen. Final report (draft b) on EUROMET
project no. 446: International comparison of magnetic flux density by means of field coil
transfer standards. Metrologia, 38(2):187–191, 2001.
185. S.A.J. Wiegers, A. den Ouden, J. Rook, J.A.A.J. Perenboom, H.H.J. ten Kate, M.D. Bird,
A. Bonito-Oliva, and J.C. Maan. Conceptual design of the 45 T hybrid magnet at the nijmegen
high field magnet laboratory. IEEE Trans. Appl. Supercond., 20(3):688–691, 2010.
186. M. Willemin, C. Rossel, J. Brugger, M.H. Despont, H. Rothuizen, P. Vettiger, J. Hofer,
and H. Keller. Piezoresistive cantilever designed for torque magnetometry. J. Appl. Phys.,
83(3):1163, 1998.
187. M. Wilson. Superconducting magnets. Clarendon Press, Oxford, 1983.
188. Wuhan High Magnetic Field Center (WHMFC), Huazhong University of Science and
Technology (HUST), Wuhan, China. http://whmfc.hust.edu.cn/english.
189. H.H. Yang, N.V. Myung, J. Yee, D.Y. Park, B.Y. Yoo, M. Schwartz, K. Nobe, and J.W. Judy.
Ferromagnetic micromechanical magnetometer. Sens. Actuators A Phys., 97-98:88–97, 2002.
190. L. Yin, J.S. Xia, N.S. Sullivan, V.S. Zapf, and A. Paduan-Filho. Magnetic susceptibility
measurements at ultra-low temperatures. J. Low Temp. Phys., 158(3-4):710–715, 2009.
191. Sangwon Yoon, Jaemin Kim, Hunju Lee, Seungyong Hahn, and Seung-Hyun Moon. 26 T 35
mm all-GdBa2 Cu3 O7−x multi-width no-insulation superconducting magnet. Supercond. Sci.
Technol., 29(4):04LT04, 2016.
192. J. Zhai, Z. Xing, S. Dong, J. Li, and D. Viehland. Magnetoelectric laminate composites: An
overview. J. Am. Ceram. Soc., 91(2):351, 2008.
193. H. Zijlstra. Experimental Methods in Magnetism: Measurement of magnetic quantities.
Number 9. 2 in Series of monographs on selected topics in solid state physics. North-Holland
Publishing Company, 1967.
194. H. Zijlstra. A vibrating reed magnetometer for microscopic particles. Rev. Sci. Instr.,
41(8):1241–1243, 1970.
195. O. Züger and D. Rugar. First images from a magnetic resonance force microscope. Appl.
Phys. Lett., 63(18):2496, 1993.

Oliver Portugall graduated at Würzburg University and SUNY


Albany before getting his PhD at the Technical University Braun-
schweig in 1992. He worked at ISSP Tokyo and Humboldt Uni-
versity Berlin before becoming deputy director of the Toulouse
pulsed field facility in 2000. Since 2021, he is again focusing on
his original activity, the generation of megagauss magnetic fields
and their application in condensed matter research.
1152 O. Portugall et al.

Steffen Krämer obtained his PhD in Nuclear Magnetic Reso-


nance (NMR) in solid-state physics at the University of Stuttgart,
Germany. He works as a senior research engineer at the Labo-
ratoire National des Champs Magnetiques Intenses in Grenoble,
France. He develops NMR instrumentation and methods for
very high magnetic fields and is responsible for the LNCMI
instrumentation and user support team.

Yurii Skourski graduated at Moscow State University in 1994,


where he also obtained his PhD in solid-state physics in 2000 after
a fellowship at the International Laboratory of High Magnetic
Fields and Low Temperatures, Wroclaw, Poland. He worked at
IFW Dresden before becoming a staff scientist at Dresden High
Magnetic Field Laboratory where he is involved in instrumenta-
tion development and support.
Material Preparation and Thin Film Growth
23
Amilcar Bedoya-Pinto , Kai Chang, Mahesh G. Samant,
and Stuart S. P. Parkin

Contents
Thin Film Growth: General Concepts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1155
Phase Transition Out of Equilibrium, Surface Kinetics and Growth Modes . . . . . . . . . . . 1155
Choice of Substrate and Role of Buffer Layer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1159
Molecular Beam Epitaxy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1164
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1164
The Ultrahigh Vacuum Environment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1165
Control of Molecular Beam Fluxes and Substrate Temperature During Growth . . . . . . . . 1167
The In Situ Monitoring of the Film Growth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1168
The Applications of MBE in Novel Functional Materials . . . . . . . . . . . . . . . . . . . . . . . . . 1169
Comparison with Other Deposition Techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1178
Pulsed Laser Deposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1179
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1179
Target Preparation and Manipulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1181
Laser . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1181
Ablation Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1182
Optics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1183
Thickness Monitoring . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1183
Substrate Heating . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1184
Selected Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1184
Magnetron Sputtering Deposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1186
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1186
Material Targets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1186

A. Bedoya-Pinto () · S. S. P. Parkin


Max Planck Institute of Microstructure Physics, Halle (Saale), Germany
e-mail: stuart.parkin@mpi-halle.mpg.de
K. Chang
Beijing Academy of Quantum Information Sciences, Beijing, China
e-mail: changkai@baqis.ac.cn
M. G. Samant
IBM Research, San Jose, CA, USA
e-mail: mgsamant@us.ibm.com

© Springer Nature Switzerland AG 2021 1153


J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_23
1154 A. Bedoya-Pinto et al.

Magnetron Sources . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1187


Magnetron Operation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1187
Chimneys and Shutters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1189
Reactive Magnetron Sputtering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1190
RF Magnetron Sputtering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1190
Growth of Alloy Films . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1191
Ion Beam Sputter Deposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1192
Ion Beam Deposition Source . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1193
Ion Beam Deposition Geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1193
Ion Beam Deposition Targets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1194
Ion Beam Deposition Source Operation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1195
Selected Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1195
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1196

Abstract

The preparation of materials in thin film form not only is of technological


importance for device applications but also grants access to study and understand
the physical properties (magnetic, optical, electronic, acoustic) of materials
approaching the two-dimensional limit. One interesting example thereof, which
is relevant for this handbook, is the occurrence of long-range magnetic order in
two dimensions, initially ruled out by theory (Hohenberg, Phys Rev 158:383–
386, 1967; Mermin, Wagner, Phys Rev Lett 17:1133–1136, 1966) but later
thoroughly revisited as the advances in thin film growth enabled the preparation
of suitable experimental testbeds (Vaz, Bland, Lauhoff, Reports Prog Phys
71:056501, 2008). Now that even the growth of innately 2D materials has
been recently achieved by ultrahigh vacuum methods (Chen et al, Science
366:983–987, 2019; Liu et al, npj 2D Mater Appl 1:1–6, 2017; Bedoya-
Pinto et al, Intrinsic 2DXY ferromagnetism in a van der Waals monolayer.
arXiv:2006.07605), the key mechanisms to stabilize magnetic order in two
dimensions have been finally established. The current level of understanding in
such a fundamental, long-debated topic would have been not possible without
the successful growth and characterization of atomically thin films on a variety
of surfaces, using the appropriate methods for a high-quality bottom-up material
synthesis.
In this chapter, we will briefly describe the most important methods for thin
film growth, such as molecular beam epitaxy (MBE), pulsed-laser deposition
(PLD), as well as magnetron and ion-beam sputtering. We will start with an
introductory section which is common to these vapor phase methods such as
the growth modes, the importance of substrate preparation, and the role of
buffer layers (Sect. “Thin Film Growth: General Concepts”.) and then describe
the characteristic features of each deposition technique, highlighting practical
aspects from the user point of view and including some representative examples
of thin film compounds grown with these methods (Sects. “Molecular Beam
Epitaxy”, “Pulsed Laser Deposition”, “Magnetron Sputtering Deposition”, and
“Ion Beam Sputter Deposition”).
23 Material Preparation and Thin Film Growth 1155

Thin Film Growth: General Concepts

The growth of thin films relies on an effective condensation of atoms from the
vapor (or liquid) phase onto a solid surface. While it may seem a straightforward
statement, the real scenario is much more complex: the atoms undergo several
competing kinetic processes when they arrive to a particular surface. These kinetic
processes – such as adsorption, diffusion, or desorption – are strongly dependent on
material parameters, but they can also be efficiently steered by adjusting external
conditions such as the atomic flux or the surface temperature. The way the atoms
arrange on the surface has been classified according to their morphology – the so-
called growth modes – and is a key factor to determine the physical properties of the
films. In most of the cases, the arrangement of the atoms at the surface can be hardly
reversed because the vapor-solid phase transition occurs out of thermal equilibrium,
such that special attention to the growth conditions needs to be paid at the very early
stage of growth. In that regard, the properties of the substrate play a decisive role,
and a right substrate choice is often the clue to develop the thin film growth of new
materials or a good starting platform to achieve unprecedented film properties. A
brief review about the concepts of vapor-solid phase transition out of equilibrium,
surface kinetics, and growth modes and the role of the substrate and buffer layers
will be described in the following sections.

Phase Transition Out of Equilibrium, Surface Kinetics and Growth


Modes

The central concept of thin film growth under high vacuum conditions is to achieve
a controlled phase transition of atomic entities (either elements or molecules) from
the vapor to the solid phase, in a thermodynamic regime out of equilibrium. For
this process to occur, the pressure of the atomic or molecular beam pbeam has to be
higher than the vapor pressure at thermal equilibrium peq of the element. From a
more general perspective, the tendency of a system to undergo a phase transition
can be described by the Gibbs free energy G = U + p V – T S, where U is the
internal energy, S is the entropy, and p,V is the pressure and the volume of the
system (in our case, the growth region), respectively. A positive balance of the Gibbs
free energy between initial and final state (G > 0) will lead to a phase transition,
while for a negative G, the system will remain in the same phase. To apply these
concepts to a thin film growth experiment, one has to assume an infinite material
reservoir providing a constant gas pressure (pbeam ) and a thermal bath providing a
constant temperature to the growth zone (Tsubstrate ), conditions which are reasonably
fulfilled in a typical experiment. In that way, a calculation of the entropy S and
enthalpy H (H = U + pV), leading to the Gibbs free energy, can be performed to
judge if the solid-state growth of certain substances is thermodynamically possible
at all. A more detailed theoretical roadmap to calculate these quantities can be found
in [1]. These considerations are particularly relevant if one aims to grow a particular
1156 A. Bedoya-Pinto et al.

element or compound which does not exist yet in thin film form. Otherwise, it is
always useful to look at existing phase diagrams.
While the first step is to assess that a vapor-solid phase transition is thermo-
dynamically accessible within reasonable conditions of temperature (Tsubstrate ) and
vapor pressures (pbeam ), it is of paramount importance for the film properties to
determine how this vapor condenses at the surface. To this end, one has to consider
the following kinetic processes experienced by the atoms once they arrive to the
surface:

• Adsorption (immobilization/lattice incorporation)


• Desorption
• Diffusion (either surface diffusion or interdiffusion/penetration)
• Aggregation (nucleation/clustering)
• Decomposition

Adsorption and desorption are opposite processes, which generally occur when
the atoms either stay at or leave the surface. Adsorption occurs when the atoms get
immobilized due to attractive potentials at the surface (chemical, electrostatic). This
is likely to occur at defect sites/dangling bonds, step edges, or electrostatic potentials
from polar substrates. One specific example of adsorption (and which is desired
in a thin film growth experiment) is when the impinging atom chemically binds
and incorporates to the surface lattice. This process can be conditioned by proper
surface functionalization of the substrate, e.g., by constructing a particular atomic
termination and dangling bond configuration which is energetically favorable for the
atom to bind. Desorption is the reverse process, when immobilized atoms (either
chemically bound or unbound) are released from the surface to the vapor phase,
usually happening after some amount of energy (mostly thermal) is given to the
system. Another very relevant case is when the atoms do not have the amount of
energy needed to leave the surface, but still enough to overcome surface potentials
and remain mobile at the surface: the diffusion regime. Surface diffusion is a
key parameter to control the growth morphology of the thin films. Under some
circumstances, the diffusion of atoms is not restricted to the surface. Especially
light elements can diffuse underneath the surface through the crystalline substrate
or underlayer (bulk diffusion), until they find a trapping potential in the solid. One
has to be aware of this phenomenon, as it may alter the properties of the substrate
or the underlayer, being the latter especially relevant when growing multilayers
or heterostructures. Aggregation occurs when an atom prefers to stick next to
another atom already residing at the surface instead of finding its own way. The
first atoms sticking on the surface are thus called nucleation centers, from which
the material growth should develop. The aggregation process can be two- or three-
dimensional: A 2D aggregation is needed for a layer-by-layer growth, whereas a 3D
aggregation will lead to the formation of islands or clusters. Last but not the least,
the decomposition process, which happens when a compound or molecule loses one
or more of its constituents, can occur both on the surface and on the vapor phase side
(thermal cracking). Decomposition is generally unwanted, as the fragmentation of
23 Material Preparation and Thin Film Growth 1157

molecules or substrate bonds releases new species that enter the kinetic scenario and
thus re-arrange the formation of phases. However, there are some unusual routes
that rely on controlled decomposition to achieve, for instance, doping or surface
templating. All these relevant processes, only briefly described here for a general
understanding and depicted schematically in Fig. 1a, are extensively discussed in
specialized books and reviews [1–3].
Since it is extremely difficult to account for a full theoretical description of
the kinetic processes due to the large number of variables, an empiric approach is
routinely applied to develop thin film growth recipes. However, a good understand-
ing of the growth kinetics and the knowledge of some material parameters (e.g.,
surface energy, cohesion forces) is very useful to start a thin film growth experiment
within a reasonable parameter window. The interplay of the aforementioned kinetic
processes will determine the growth mode of the film, which can range from a
two-dimensional, layer-by-layer growth to a fully three-dimensional (island) or
columnar growth. These are summarized schematically in Fig. 1b, along with the
morphology evolution during the initial stage of growth. A layer-by-layer growth,
also known as Frank-van der Merve growth, generally sets in when the atoms
have a stronger affinity to stick to the surface than to aggregate to each other. In
the absence of nucleation centers, i.e., in a perfectly flat and defect-free surface,
the atoms would find available spots on the surface until the first monolayer is
completed. But in reality, the first atoms reaching the surface will get pinned at
defects, step edges, or any type of attractive surface potentials, from which the
nucleation and crystal growth process will start, provided that the atoms (either
within the same of different species) have a natural affinity to bind and form a
(crystalline) structure. Important for the Frank-van der Merve growth regime is
that, regardless of the number of nucleation centers, the growth of the second layer
does not start until the full surface coverage is accomplished by the first layer.
Diffusion plays here an important role, as the atoms might move sideways until
they find their energetically most favorable position. In cases where the substrate
surface present a regular terrace structure (originated by a slight miscut angle during
preparation, see Sect. “Choice of Substrate and Role of Buffer Layer”. for details),

Fig. 1 (a) Kinetic processes occurring in a solid surface upon exposure to gaseous molecular
beams. (b) Schematic representation of the main growth modes and the arrangement of the atoms
at the early stage of growth. (Taken from [4])
1158 A. Bedoya-Pinto et al.

the Frank-van der Merve growth manifests in the homogeneous atomic coverage
of each terrace, provided that the diffusion length of the atoms in the surface is
longer than the terrace width. This special case, favoring a layer-by-layer growth
over regular step terraces, is often denoted as step-flow growth. In particular, where
the substrate step terraces have monoatomic height similar to the lattice parameters
of the growing crystal, the subsequent layers will accommodate on top of each
other without building stacking faults and yielding a high crystalline order. Thus,
the step-flow layer-by-layer growth is one of most sought-after conditions for a
defect-free and high-quality epitaxy on substrates displaying a finite miscut angle.
On the other hand, a layer-by-layer growth is disrupted if the cohesion forces
between the atoms are larger than the affinity for them to bind to the surface.
This will lead to a volumetric (3D) aggregation of atoms at distinct nucleation
centers of the surface, denoted as Volmer-Weber or island growth. Depending on
the interplay between diffusion and aggregation, these islands might coalesce at a
particular film thickness. Even if the islands merge with each other, the Volmer-
Weber growth mode generally leads to a high surface roughness and large structural
mosaicity, which is usually unwanted for thin film applications. Interestingly, there
is a special case within the Volmer-Weber regime when the coalescence of islands
is energetically not favorable, such that the growth evolves perpendicular to the
surface (columnar growth). This directional growth regime has been proven useful
to achieve the bottom-up growth of nanowires/nanowhiskers of high crystalline
quality [5, 6], as each column will not experience much strain forces during the
directional growth. The diameter of the nanowires can be controlled by (artificially)
setting the density of nucleation centers on the surface. Last but not the least, a
combination of layer-by-layer and island growth can also occur during a growth
process, denoted as Stranski-Krastanov (mixed type) growth. It happens usually
when the growth starts in a layer-by-layer fashion, but later turns to aggregation
and three-dimensional growth, as exemplified in Fig. 1b. One reason is simply that
when the growth starts, the undergoing kinetic processes happen at the substrate
surface, while the situation changes as the growth evolves and the substrate is fully
covered by the material under study, constituting a “new” surface for the impinging
atoms. One route to overcome this is to change the growth parameters (e.g.,
temperature, atomic flux) after the first layer(s) covers the substrate surface, such
that conditions for layer-by-layer growth can be re-established (two-step growth).
In the case of homoepitaxy (a thin film grown on a substrate of the same kind),
although less common, a Stranski-Krastanov growth may set in if the surface energy
or the lattice parameters experience a change as the thickness of the grown layer
increases.
The knowledge of the kinetic processes and resulting growth modes, their
identification, and the correlation with the growth parameters lie at the heart of a thin
film growth experiment (typically, the growth modes and its resulting morphology
are assessed via in situ methods such as electron diffraction or surface topography
scanning probes). In this regard, a decisive role is played by the substrate surface,
as in many practical cases, a bad choice of the substrate or a deficient surface
preparation cannot be compensated by adjusting the growth parameters. Therefore,
23 Material Preparation and Thin Film Growth 1159

the next subsection will be devoted to the role of substrates; and in the cases where
optimal substrates cannot be found or prepared, the preparation of a buffer layer as
a n alternative solution will be underlined.

Choice of Substrate and Role of Buffer Layer

One important aspect for a successful epitaxial growth is the choice and preparation
of the substrates. As a rule of thumb, the following criteria need to be taken into
account:

• Lattice matching between substrate and film (ideally within 3% mismatch)


• High temperature stability and robustness of the substrate surface
• Low chemical reactivity
• Similar thermal expansion coefficient of substrate and film
• Atomically flat surfaces attainable via cleaning treatment recipes
• Low amount of structural defects (e.g., dislocations, stacking faults)

It should be noted that for many cases, it is not so easy to find substrates that
fulfill all these requirements. One can still achieve epitaxial growth even if not all
these conditions are strictly met, but one has to be aware that additional actions need
to be implemented, adding more complexity to the substrate preparation process.
For instance, the problem of lattice matching between substrate and film can be
overcome, to a certain extent, by the growth of an intermediate buffer layer. On
the other hand, the temperature stability and chemical reactivity of the substrate
are more critical parameters, as the substrate surface might undergo irreversible
structural changes upon heating or exposure to particular atomic fluxes. The
matching of thermal expansion coefficients is important to avoid inhomogeneous
compression/expansion or substrate and film upon heating/cooling cycles, which
might lead to cracking and rippling of the grown layers. Last but not the least, the
preparation of a substrate with an atomically flat surface (via chemical or thermal
treatment) reduces the diffusion barriers driven by topography and thus facilitates a
layer-by-layer growth mode.
Typical substrates that are used for epitaxy (besides Si-wafers, which are one
of the most popular ones due to CMOS compatibility) are oxides and nitrides.
They have the advantage to be, in general, chemically inert due to a strong ionic
and/or covalent bonding and thermally stable up to relatively high temperatures.
In addition, there is a vast knowledge about surface preparation recipes on these
compounds [7], to get rid of surface contamination as well as to achieve an
atomically flat topography. Figure 2 shows an example of how chemical cleaning
and high-temperature annealing procedures impact the surface topography of
commercially available substrates, such as Al2 O3 (0001). While a modified RCA
(named after Radio Corporation of America, originally developed for Si-wafers)
cleaning treatment is able to get rid of surface adsorbates (Fig. 2b), an additional
high-temperature annealing (1150 ◦ C) under O2 atmosphere results in an ordered,
1160 A. Bedoya-Pinto et al.

Fig. 2 Atomic-force microscopy images of Al2 O3 (0001) substrates (a) as received from the
commercial supplier, (b) after modified RCA cleaning of the substrate, and (c) after RCA cleaning
and high-temperature annealing (1150 ◦ C) in a furnace under oxygen atmosphere. The chemical
cleaning step is efficient in removing all the impurities on the substrate surface, and atomically thin
step terraces are visible after the annealing procedure

atomically flat surface, featuring step terraces as thin as one unit cell of the Al2 O3
(0001) crystal (Fig. 2c). This example shows how chemical and heating treatment
recipes are key to produce a substrate surface with ideal conditions to initiate a
controlled layer-by-layer, step-flow growth.
Most of the commercially available substrates are produced as bulk crystals with
extremely large structural coherence and a low number of defects and dislocations;
and then they are mechanically cut to a particular size and surface-treated via
chemical-mechanical polishing (CMP) procedures. Both of these characteristics
from surface-finished bulk substrates (low defect density and surface flatness) are
favorable conditions for thin film epitaxy. However, there are cases where bulk
crystal synthesis of substrates is extremely difficult and can be only performed under
extreme conditions. For example, gallium nitride (GaN), a very common compound
used in III–V semiconductor applications and a popular substrate for nitride
epitaxial growth, can only be synthesized at 1700 ◦ C and at pressures (20 kbar)
to overcome the low solubility of N in the molten Ga liquid [8]. An alternative to
bulk crystal substrates that has been developed is the fabrication of “freestanding”
substrates, which result from the epitaxial growth of a given material on a foreign
substrate (heteroepitaxy) and its subsequent detaching from it. In the case of GaN,
freestanding substrates are usually prepared by epitaxial growth of a very thick
GaN layer (>100 μm) on sapphire or SiC substrates, followed by a delamination of
the GaN layers from the substrates via a laser-induced separation method [9]. One
important issue to consider here is that the quality of the freestanding thick layer is
determined from the growth procedure on the initial substrate. Thicker freestanding
substrates have usually superior structural quality, since in heteroepitaxial systems,
the number of dislocations reduces with film thickness; but the production costs are
also higher. Another – less expensive – alternative for epitaxial growth that does not
involve delamination is to use the layers still attached to their foreign substrates, also
called heteroepitaxial “templates. ” The structural quality and surface morphology
23 Material Preparation and Thin Film Growth 1161

Fig. 3 Surface topography of different GaN (0001) substrates. (a) GaN (0001)/Al2 O3 (0001)
template grown by MOCVD, showing atomically sharp terraces and a low number of dislocations
(black areas). (b) Bulk GaN synthesized by high-temperature and high-pressure methods. Although
it presents a low roughness, no atomic steps are visible. (c) GaN (0001) template grown on Si (111),
showing the highest roughness and the largest amount of dislocation densities. (d) Rocking curve
of the asymmetric (104) GaN reflection, showing that the bulk crystal has the lowest full width at
half maximum and thus the best crystallinity. (Courtesy: Christian Zube, University of Göttingen)

of these templates vary depending on the lattice matching to the substrate and the
thickness and the thermal expansion coefficients. Figure 3 shows an example on
how different substrate production strategies affect their surface morphology and
structural quality. While the bulk GaN (0001) substrates synthesized with high
pressure and temperature methods display the lowest roughness (Fig. 3b) and best
rocking curve widths, a measure of the crystalline coherence (Fig. 3d), the surface
does not present atomic step terraces. On the other hand, the GaN(0001) “templates”
grown on Al2 O3 (0001) and Si (111) (Fig. 3a, c) both present a surface featuring
step terraces but display wider rocking curve widths indicative of more structural
defects. Comparing only the two heteroepitaxial templates, the GaN (0001) grown
on Si (111) exhibits a larger structural defect density (the widest rocking curve
width, Fig. 3d), which, in turn, manifests also in an irregular surface step terrace
pattern and a higher surface roughness (Fig. 3c). This difference is attributed to
the large lattice mismatch (17%) between Si (111) and GaN (0001) and dissimilar
thermal expansion coefficients, and although this has been partially mitigated by the
insertion of AlN/GaN buffer layer superlattices in up-to-date industrial processes,
the substrate quality is still moderate. The GaN (0001) template on Al2 O3 (0001), on
the other hand, has an overall good performance in terms of surface roughness, step
terrace homogeneity, and crystalline quality, added to the advantages of low cost
compared to freestanding substrates. These are the usual benchmarks to consider
for a proper choice of substrate provider.
In the last years, the accelerated discovery of materials with exotic properties
gave rise to an ever-increasing demand to grow thin films of complex crystalline
structures and atomic compositions, for which a substrate fulfilling the optimum
requirements – listed at the beginning of Sect. “Choice of Substrate and Role of
Buffer Layer” – is difficult to find. One of the biggest problems in this regard is
the structural lattice mismatch of the material structure to the available substrates.
However, there are some strategies to overcome this limitation, such as:
1162 A. Bedoya-Pinto et al.

(i) To find a geometrical overlap of multiple surface unit cells into a single one
(e.g., multiple cubes into one rectangle)
(ii) To find a lattice matching by allowing the rotation of substrate and film surface
unit cells under specific angles
(iii) To grow one (or more) buffer layer(s) with lattice parameters in between the
values of substrate and film, effectively reducing the lattice mismatch

All these strategies are not mutually exclusive, and they can be implemented in
the same growth procedure if required. In the following, the rational use of these
strategies will be exemplified with the help of two novel topological materials, WP2
and NbP, each of them having a peculiar crystal structure which makes it difficult
to lattice-match with commercially available substrates. WP2 has an orthorhombic
structure (non-symmorphic Cmc21 space group) with very dissimilar a (3.17 Å),
b (11.16 Å), and c (4.97 Å) lattice parameters, but which is highly interesting
to explore in thin film form, due to its exotic electron transport properties which
resemble the ones of a hydrodynamic fluid [10]. Among insulating substrates, which
are desirable to explore the intrinsic transport properties of the films, one of the very
few options is to use AlN (in the wurtzite structure), but using the (1–100) plane,
unlike most of the epitaxial processes which rely on the hexagonal (0001) plane. In
this configuration, the a and c lattice parameters of AlN match well with the ones
of WP2 (mismatch of 1.7% and −0.16%, respectively), as depicted schematically
in Fig. 4a. This example shows that a careful search on non-conventional crystal
faces among the large substrate database is often the key to find the solution of an
epitaxial growth problem. However, this provides only a solution for an epitaxial
growth along the b-axis of WP2 (a–c plane on the surface). As many physical
properties depend on the crystalline orientation (for instance, a large crystalline
anisotropy of the magnetoresistance has been measured in WP2 bulk crystals [11]),
it is often required to pursue the epitaxial growth along different orientations. For
an orthorhombic structure with dissimilar lattice parameters like WP2 , this can be
a real challenge. A strategy to grow WP2 along c-direction (which means a basal
plane of a = 3.17 Å and b = 11.16 Å) is found by the geometrical fitting of multiple
surface unit cells onto a given structure. As depicted in Fig. 4b, this condition can
be achieved by fitting three rectangular a–b units into a non-integer multiple of a
hexagonal surface lattice, such as the GaN (0001) or NdGaO3 (011) plane. This
yields a lattice mismatch of 0.8% and 2.1% in a and b of WP2 , respectively. Such
an unusual geometrical lattice fit like WP2 on GaN is usually more challenging to
achieve in a real growth experiment; however, they shall be mentioned for the sake
of completeness and to encourage the reader to think also about these possibilities
for a lattice-matched epitaxial growth.
A second example, where the lattice mismatch is reduced by the rotation of the
unit cell in the surface plane and a subsequent growth of a buffer layer, is highlighted
by the growth of NbP thin films on insulating oxides. NbP, a recently discovered
topological semimetal with ultrahigh electronic mobility [12], crystallizes in a
tetragonal structure (space group I41 md) with lattice parameters a = 3.34 Å and
c = 11.37 Å, as depicted in Fig. 5a. Although it features a square lattice in the
23 Material Preparation and Thin Film Growth 1163

Fig. 4 Lattice-matched epitaxial growth strategies for orthorhombic WP2 (a = 3.17 Å,


b = 11.16 Å, c = 4.97 Å). (a) For an epitaxial growth along b-axis, the m-plane (1–100) of wurtzite
AlN turns out to be suitable. (b) As for c-axis-oriented growth, the situation is more complicated
due to the highly dissimilar a and b lattice parameters of the basal plane. A proposal based on a
geometrical fitting of rectangular (a–b plane of WP2 ) on hexagonal lattices (such as GaN (0001)
or NdGaO3 (011)) is highlighted

basal a-a plane – and thus a less complicated case than orthorhombic WP2 – the
available palette of lattice-matched cubic and perovskite substrates is scarce. For
square surface lattices, one intuitive strategy to account for a lattice matching
is to consider a rotation of 45◦ between substrate and growing film. Hence, by
choosing either MgO (a = 4.29 Å) or rutile TiO2 (a = 4.59 Å) and rotating it
45◦ to get a lattice distance of 3.02 Å and 3.24 Å in (110) direction, respectively
(Fig. 5b), one approaches the lattice constant of NbP. While for the case of TiO2 ,
the mismatch to NbP (3.34 Å) is still reasonable (<3%) to proceed with an epitaxial
growth experiment, the growth of NbP on MgO substrates will require additional
efforts to improve the lattice matching: the growth of a buffer layer. Cubic Nb
(100), with a lattice parameter of a = 3.30 Å, lying in between substrate and film,
appears as a good buffer layer candidate: in fact, high-quality epitaxial films of
NbP have been grown using this strategy [13]. The role of the buffer layer, or
sometimes called sacrificial layer, is to reduce the lattice mismatch at expenses of
sustaining a large amount of structural defects, thus improving the situation for the
subsequent growth of the layer of interest. The defect density typically reduces
with increasing buffer layer thickness, so that it has to be individually adjusted
for each epitaxial growth process – the initial mismatch is also different in each
case. In situ characterization techniques, such as reflection high-energy electron
diffraction (RHEED) and scanning tunneling microscopy (STM), are of great help
to judge whether the surface quality of the buffer layer of a particular thickness
is good enough for the subsequent epitaxy of the layer of interest. Figure 5c
describes the growth evolution of NbP (001) on MgO (001) using a Nb (001)
1164 A. Bedoya-Pinto et al.

Fig. 5 (a) Tetragonal, non-centrosymmetric crystal structure of NbP. (b) Surface unit cell of MgO
(001), highlighting the lattice distances a100 and a110 upon 45◦ in-plane rotation that can be used
for epitaxial growth. (c) Schematic description of a growth experiment of NbP on MgO using a Nb
buffer layer, along with in situ RHEED patterns of the growth evolution. The 45◦ in-plane rotation
between MgO and Nb (NbP) can be visualized in the top view diagram on the right, yielding an
in-plane epitaxial relationship MgO(100)//NbP(110). (d) Transmission electron microscopy image
of the grown MgO/Nb/NbP stack, where the structural quality of film and buffer layer can be
visualized. Panels (c) and (d) are adapted from [13]

buffer layer, highlighting the crystal structures and epitaxial arrangement in cross
section and top view, along with in situ electron diffraction (RHEED) patterns of
the crystal structures at the various growth stages. Several important features shall
be underlined here: (i) the in-plane rotational growth strategy between MgO and Nb
(NbP) lattices has been successfully implemented in the experiment, as evidenced
by the coincident RHEED patterns of MgO (100) and Nb (110), (ii) a small deviation
of the RHEED streak positions between MgO (100) and Nb (110) – indicative
of a slightly different lattice parameter – hints that the Nb buffer layer surface is
not strained but has relaxed to its natural lattice constant, and (iii) the Nb buffer
layer presents broad, modulated RHEED streaks, whereas the NbP layer grown on
top presents a more sharp and streaky RHEED pattern, indicative of an improved
crystalline quality and surface flatness. To provide the reader with a real-space
visualization of the growth outcome, Fig. 5d shows a cross-sectional transmission
electron microscopy (TEM) image of the MgO/Nb/NbP stack, where a flat, highly
ordered NbP layer can be observed, at expenses of a more structurally defective
Nb buffer layer. Hence, this particular experimental example is very instructive
both from the point of view of growth strategy and choice of substrate (45◦ in-
plane rotated epitaxy on a commercially available, chemically robust substrate)
and the role of buffer layer (“buffering” the structural damage arising from lattice
mismatch).

Molecular Beam Epitaxy

Introduction

Molecular beam epitaxy (MBE) is a process in which localized atomic or molecular


beams are supplied in ultrahigh vacuum (UHV) environment as sources for the
23 Material Preparation and Thin Film Growth 1165

epitaxial growth of materials on a crystalline substrate [14]. There are several


essential features in an MBE process: (i) the mean free path of the molecular beams
is comparable or larger than the distance between the effusion cells and the substrate,
i.e., ballistic molecular beams, (ii) the grown material and the substrate usually share
the same lattice symmetry and similar lattice parameters at their interface, and (iii)
the substrate temperature is usually much lower than that in single crystal synthesis,
far away from thermal equilibrium. The first feature allows abrupt switching of
molecular beams by mechanical shutters; the second feature guarantees large-scale
single-crystalline growth; and the third feature prevents the atoms from diffusing
across the interface. These features make MBE a powerful tool for the growth of thin
films or nanostructures with low defect concentration, as well as for the fabrication
of atomically sharp heterostructures and superlattices.
The history of MBE greatly overlaps with the development of semiconductor
industry [15] [16]. Since its invention in the 1960s, it has been widely applied to the
growth of III–V, II–IV, IV–VI, IV–IV, and oxide semiconductors. The most famous
application of MBE is probably the growth of III–V semiconductor heterojunctions
hosting two-dimensional electron gases (2DEG) with extremely high mobility,
which leads to the discovery of integer and fractional quantum Hall effects, awarded
with the Nobel Prizes in 1985 and 1998 [17, 18]. There are two more Nobel Prizes
that have directly benefited from the development of MBE: the discovery of electron
tunneling phenomenon (1973) [19] and the fabrication of modern heterostructure
electronics (2000) [20]. In the twenty-first century, MBE is playing a more and more
important role in the discovery of new functional materials, such as topological
insulators, high-temperature superconductors, and two-dimensional (2D) van der
Waals materials. In this section, we will briefly introduce the MBE environment
and growth control in Sects. “The Ultrahigh Vacuum Environment”, “Control of
Molecular Beam Fluxes and Substrate Temperature During Growth”, and “The In
Situ Monitoring of the Film Growth”; review the applications of MBE in Sect.
“The Applications of MBE in Novel Functional Materials”, putting emphasis on
the recently developed materials; and discuss the advantages and limitations as well
as the future perspectives of the MBE technique in Sect. “Comparison with Other
Deposition Techniques”.
The schematic diagram of a typical MBE chamber is shown in Fig. 6. The core of
the MBE technology mainly includes three aspects: (i) the acquisition of a suitable
ultrahigh vacuum (UHV) environment, (ii) the precise control of molecular beam
fluxes and substrate temperature, and (iii) the in situ monitoring of the film growth.
We will discuss how each aspect contributes to the growth of high-quality films in
the following sections.

The Ultrahigh Vacuum Environment

Thin films grown by MBE generally feature an atomic-precise control of the


thickness and high crystalline quality, which demand a relatively slow growth
rate. The requirement of ballistic molecular beams also limits the growth rate. For
1166 A. Bedoya-Pinto et al.

Fig. 6 Schematic diagram of an MBE chamber

instance, the effusion cells usually work at the vapor pressures of 10−4 –10−6 Torr,
which correspond to molecular mean free paths from several centimeters to several
meters, comparable to the distance between the effusion cells and the substrate
(∼20 cm). A typical growth rate of III–V semiconductors is 1 monolayer (ML)
per second, or ∼ 1 μm/h [12]. In some latest applications of ultrathin film growth,
such as topological insulators [21] and transition metal dichalcogenides [22], the
growth rate can be as low as 1 ML/min or even 1 ML/h. An extremely clean
environment is required to avoid unintentional doping at such low growth rates.
For example, the number of O2 molecules impinging on a surface per second is
 1/2
NA
given by = p 2π Mk B T , in which p is the partial pressure of O2 , M is the molar
mass of O2 , NA is the Avogadro’s constant, kB is Boltzmann’s constant, and T is
23 Material Preparation and Thin Film Growth 1167

the absolute temperature. When growing Bi (111) films at the rate of 1 ML/s in a
vacuum environment with an oxygen partial pressure of 10−8 Torr, assuming 10%
of the oxygen molecules react with bismuth, then the impurity level in the bismuth
film is 0.07%, which is much higher than the one found in commercially available
bulk materials (0.0001% – 0.001%). Therefore, most MBE applications must be
carried out in ultrahigh vacuum, typically 10−9 – 10−12 Torr.
MBE chambers are designed to maintain UHV environments during the depo-
sition. The pumping system of an MBE chamber usually combines rotary, turbo
molecular, ion getter, and titanium sublimation pumps. Cryogenic pumps are often
installed in the medium- to large-scale MBE systems. The chamber is built with non-
magnetic stainless steel, usually with full metallic sealing. In order to quickly pump
out the water molecules, which adhere to the chamber walls and desorb very slowly
at room temperature, the chamber is baked at 150 – 200 ◦ C for several tens of hours
after every exposure to atmospheric conditions (e.g., for maintenance). Besides,
during the growth, liquid nitrogen or cold gas flows through the cryoshrouds inside
the chamber. The cryoshrouds serve as another cryopump, absorbing the residual gas
molecules and further reducing the pressure. All the heating elements, including the
sample stage and the effusion cells, are water-cooled or shielded in order to avoid
strong outgassing.
The substrates are introduced through a load-lock, which is a small chamber that
can be quickly pumped to 10−7 –10−8 Torr before transferring the substrates into the
growth chamber. In some MBE systems that are highly sensitive to impurities, there
is an additional substrate preparation chamber between the load-lock and growth
chamber. Those “dirty” jobs, such as substrate degassing and annealing, capping
layer growth, etc., can be carried out in the preparation chamber.

Control of Molecular Beam Fluxes and Substrate Temperature


During Growth

The molecular beam fluxes can be obtained from the pseudo-Knudsen cells hosting
ultrapure solid, liquid, or gas sources. A Knudsen cell is a closed crucible with a
small orifice, through which the vapor of the material heated in the crucible leaks
out as a molecular beam. The beam flux from such a cell is solely determined by the
vapor pressure of the material, and the vapor pressure can be easily controlled by the
temperature of the crucible. In order to get higher molecular beam flux, the realistic
effusion cells used for MBE usually have large orifices. The vapor pressure required
for MBE growth is typically 10−4 –10−6 Torr. For the materials that can reach the
required vapor pressure at 100 – 2000 ◦ C, normal effusion cells with crucibles
(usually made of hBN, Al2 O3 , graphite, W, or Ta) and filament radiation heating can
be used. For those materials that require temperatures higher than 2000 ◦ C, electron
beam evaporators must be applied. The filaments in the effusion cells are connected
to PID control loops in order to stabilize the temperature. On the other hand, to
handle the gas sources, or the solid or liquid sources that already have high vapor
pressure at room temperature, one should use effusion cells with leak valves. All
1168 A. Bedoya-Pinto et al.

types of effusion cells are equipped with computer-controlled mechanical shutters,


which can switch on or off the molecular beams abruptly.
As the MBE technology develops, many specially designed effusion cells that
are dedicated to certain kinds of materials appeared. For example, in order to
reduce the danger of cracking, double-walled crucibles are used for the metals
that melt at evaporation temperatures (Al, In, Bi, etc.). Oxide MBE systems are
usually equipped with ozone delivery systems, since ozone has better reactivity than
oxygen. Some group V and VI elements (P, As, Se, Te, etc.) evaporate as tetramer
molecules. A cracker stage, which is a hot tube in front of the crucible, can break
part of these tetramers into dimers, thus increasing the reactivity. The molecular
beam fluxes can be in situ calibrated and monitored by quartz microbalances or
ionization gauges.
As for the substrate temperature, inch-scale wafers can be heated up to several
hundred degree Celsius by a radiation heater on the sample stage. Smaller substrates
can be heated up to 1200 – 1400 ◦ C using electron beams, or by passing direct
current through the substrate if it is conductive. The temperature of the substrate
is monitored by thermocouples, thermistors, or infrared pyrometers. When large
wafers are used, the sample stage keeps rotating in order to maintain the distribution
of the heating and deposition uniform. Some sample stages also integrate liquid
nitrogen cooling function for low-temperature growth.

The In Situ Monitoring of the Film Growth

The most important in situ and real-time monitoring tool is reflection high-
energy electron diffraction (RHEED), through which the lattice parameters and
reconstructions of the sample surface can be determined [23]. Especially, when
the growth is layer-by-layer (Frank-van der Merve growth), the intensity of the
diffractive pattern oscillates with time (Fig. 7). The RHEED pattern intensity
reaches a maximum at each time a new complete layer is finished, by when the
surface has fewest steps thus smoothest. Since the growth rate and surface condition
can be precisely determined from the RHEED oscillation, a technique named phase-
locked epitaxy (PLE) was developed for the growth of superlattices with atomically
sharp interfaces [15]. In PLE, the shutters only switch molecular beams when the
RHEED pattern intensity reaches a maximum; thus the intermixing at the interface
is minimized.
Since the MBE chamber is compatible with most surface analysis techniques
working in UHV, many other in situ analysis tools can be integrated with MBE, such
as X-ray photoemission spectroscopy (XPS), Auger electron spectroscopy (AES),
low-energy electron diffraction (LEED), ultraviolet photoemission spectroscopy
(UPS), and secondary ion mass spectrometry (SIMS). These techniques provide the
ability of both lattice structure and chemical composition analysis. In recent years,
many MBE systems are also combined with powerful real-space and momentum-
space imaging techniques, such as scanning tunneling microscopy (STM) and
angle-resolved photoemission spectroscopy (ARPES).
23 Material Preparation and Thin Film Growth 1169

Fig. 7 (a) The RHEEDpattern of Si(111) substrate with 7 × 7 reconstruction. Incidence azimuth:
[112]. (b) The RHEED pattern of 4 QL (1 QL = 5 atomic layers) thick Bi2 Te3 films grown on the
Si substrate. (c) The oscillation of the intensity of the zeroth-order reflected electron beam, with
different growth rates [21]. (With kind permission from Wiley)

The Applications of MBE in Novel Functional Materials

As mentioned in the first section, MBE has been very fruitful in the growth of
semiconductor thin films and heterostructures. There has been an ample literature
spectrum reviewing the contribution of MBE in these fields. Limited by the length
of this chapter, we will focus on the MBE growth of novel functional materials that
developed in the recent decade. For the MBE growth of traditional semiconductors,
we recommend the readers to refer to the literature [14–16], [23–26].
It is worth mentioning a core principle for the stoichiometric growth of not
only traditional semiconductors but also most of the novel materials – the “three-
temperature method. ” This method typically applies to binary compounds but can
also be extended to the ternary ones. Take a binary AB film, which is grown by the
co-deposition of A (usually a metal) and B (usually a nonmetal) molecular beams,
as an example: A is a low-vapor-pressure element that evaporates at the temperature
TA , while B is a high-vapor-pressure one that requires TB < TA to evaporate. At
a certain substrate temperature TS , B molecules desorb much faster than A from
the surface of the substrate. Therefore, it is recommended to choose a substrate
temperature TB < TS < TA , at which only the B molecules that bond with A stay
at the surface. Excessive B flux is provided to achieve stoichiometric growth, and
the growth rate is controlled by the A flux. For ternary compounds A1 A2 B, one
just needs to supply excessive B flux and control the flux ratio between A1 and A2 ,
thus greatly simplifying the optimization of growth parameters. It is easy to realize
that the three-temperature method is suitable for the growth of the most B-rich
compound on the A–B phase diagram, which is also true for most of high-quality
MBE-grown materials, from the traditional III–V, II–VI, and IV–VI semiconductors
to the novel functional materials that will be introduced below.

Two-Dimensional Magnetic Materials


In spirit of this handbook, the first material class that will be discussed is
the one of layered magnetic materials, which have experienced an explosion
of interest after their successful fabrication in the truly two-dimensional mono-
layer limit by mechanical exfoliation and encapsulation of atomically thin flakes
1170 A. Bedoya-Pinto et al.

[27, 28]. Importantly, recent efforts on molecular beam epitaxy of these magnetic
layered materials are currently bringing new insights on the fundamentals of
low-dimensional magnetism, such as the direct observation of stacking-dependent
magnetism in atomically thin CrBr3 [29] or the realization of a nearly ideal 2DXY
magnetic system in monolayer CrCl3 [30]. In this context, it is worth mentioning
that a vast number of MBE growth experiments of classical 3d magnets (such as
Fe, Co, and Ni) on highly crystalline substrates have been pursued down to the
monolayer limit (a comprehensive review can be found in Ref. [31]). However, in
all these highly crystalline systems, the coupling to the substrate is so strong that the
system can be hardly regarded as two-dimensional. Layered materials, on the other
hand, have a higher affinity to build two-dimensional structures than to bind to the
underlying substrate. In particular, when they are grown onto substrates with low
surface energy, such as graphene [30] or highly ordered pyrolytic graphite (HOPG)
[29], they can attain large two-dimensional grain sizes, chemically decoupled from
the substrate by a van der Waals gap. Despite the weak coupling, the choice of the
substrate will have an impact on the growth of van der Waals materials. An example
is given in Fig. 8, where a CrCl3 monolayer, a recently discovered 2D-XY magnet,
is grown on graphene/6H-SiC (0001) and directly on 6H-SiC (0001) substrates. As

Fig. 8 (a) Cross-sectional view of a CrCl3 monolayer grown on graphene/6H-SiC (0001). (b, c)
Scanning tunneling microscopy images and RHEED diffraction patterns of a CrCl3 monolayer
grown on graphene/6H-SiC and directly on 6H-SiC(0001) substrates, showing the influence on the
substrate on the in-plane orientation of crystalline domains. Twisted in-plane domains are attained
for the growth on graphene as evidenced by the dissimilar moiré patterns, while on 6H-SiC, the
CrCl3 layer grows in a single-crystalline fashion, following the in-plane symmetry of the substrate.
(d–f) Growth of Fe3 GeTe2 on Al2 O3 (0001). A layer by-layer growth is evidenced by streaky
RHEED patterns and intensity oscillations over time (e), and the epitaxial orientation is visualized
by X-ray diffraction spectra (f). (Panels (d–f) are taken from [32])
23 Material Preparation and Thin Film Growth 1171

expected, the CrCl3 grows in a (0001), i.e., c-axis orientation in both cases, but
the differences rely on the in-plane epitaxial relationships: while CrCl3 present in-
plane twisted rotational domains (as evidenced in the multiple streaks appearing
in the RHEED pattern and diverse moiré patterns on different grains observed by
scanning tunneling microscopy (STM); see Fig. 8b), the growth of CrCl3 on 6H-SiC
results in a single-crystalline orientation, the same as the underlying 6H-SiC (0001)
surface, characterized by a single-streak RHEED diffraction pattern (Fig. 8c). CrCl3
evaporates as a molecule, such that a single source is used for the MBE growth,
and the growth conditions are governed by the flux and substrate temperature
[30]. Another two-dimensional ferromagnet which has been successfully grown by
MBE is the ternary compound Fe3 GeTe2 , both on Al2 O3 (0001) and GaAs (111)
substrates [32]. Here, the growth has been carried out using the previously described
generalized three-temperature method: high-purity Fe (99.99%), Ge (99.999%), and
Te (99.999%) were co-evaporated from standard Knudsen cells at the temperature
of 1165 ◦ C, 1020 ◦ C, and 285 ◦ C, respectively [32], and the optimized substrate
temperature (340 ◦ C) lies in between the low-vapor pressure materials Fe and
Ge and the high-vapor pressure one (Te). Figure 8 (e, f) shows the structural
characterization of the films by in situ RHEED and ex situ X-ray diffraction,
evidencing epitaxial, layer-by-layer growth.

Topological Insulators
Topological insulators (TI) are a category of materials in which the electronic
structure in their k-space have a nontrivial symmetry-protected topological order.
Their electronic band structure features an inversed band sequence and a spin-orbit
interaction opened band gap. Although they have a bulk band gap like ordinary
insulators, there are symmetry-protected surface or edge states that are metallic
and contain massless Dirac fermions (linear E-k dispersion) [33]. Time-reversal
symmetry-protected 2D TI was first predicted in quantum spin Hall (QSH) systems
[34–36], and later Bi2 Se3 , Bi2 Te3 , and Sb2 Te3 were predicted to be 3D TIs with
Dirac surface states [37]. MBE has pioneered the early exploration of topological
insulators. The QSH edge states were discovered in HgTe/Hg0.3 Cr0.7 Te quantum
wells just 1 year after the prediction (Fig. 9a, b) [38]. The Dirac surface states were
also quickly confirmed from STM and ARPES experiments on the MBE-grown
films (Fig. 3c, d) [20], [39–41].
The 3D TIs are especially suitable for MBE growth because of their van der
Waals layered structure. The low interaction between the 3D TIs and the substrates
makes the requirement of lattice matching much weaker and hence broadly expands
the selection of substrates. Since 3D TIs can be grown on almost any atomically flat
surfaces, different choices of substrates can introduce different functions to the TI
films. For example, various TI/superconductor heterojunctions are grown in order to
search for Majorana bound states (MBS), which is an important candidate conceived
for quantum computation [42]. Such heterojunctions include Bi2 Se3 /NbSe2 [43],
Bi2 Te3 /NbSe2 [44], Bi2 Se3 /Bi2 Sr2 CaCu2 O8+δ [45], Bi2 Se3 /Nb (Nb as capping
layer) [46], etc. Another usage of the substrate is gating dielectric, such as
SrTiO3 (111), which has a high dielectric constant at low temperature. Quantum
1172 A. Bedoya-Pinto et al.

Fig. 9 (a) A schematic diagram of a typical HgTe quantum well [39] (With kind permission
from Wiley). (b) The quantized conductance in 7.3-nm-wide HgTe quantum wells, in which band
inversion happens (curves III and IV). From [38]. (Reprinted with permission from AAAS). (c)
STM topography (0.5 μm × 0.5 μm) image of an 80-nm-thick Bi2 Te3 film grown on Si(111)
surface. Inset, atom resolved STM image (2.5 nm × 2.5 nm). (d) Dirac surface states on the Bi2 Te3
film with linear band dispersion, resolved by in situ ARPES [20]

anomalous Hall effect (QAHE) was realized on the Cr0.15 (Bi0.1 Sb0.9 )1.85 Te3 films
grown on SrTiO3 (111), in which the bulk carriers are depleted by back-gate tuning
through SrTiO3 ; thus, the quantum conductance through the QAHE edge states
is revealed [47]. In GaAs/QAHE TI/superconductor heterostructures, half-integer
quantized conductance plateaus were observed, which is probably an evidence of
Majorana fermion modes [48]. MBE grown TIs are also promising for spintronic
applications: in Bi2 Se3 /permalloy and Bi1−x Sbx /MnGa heterostructure, spin Hall
angles as large as 3.5 and 52 are observed, much larger than those in elemental
heavy metals (< 1) [49, 50].
As the topological material family expands quickly in recent years, MBE is also
applied to the growth of topological materials beyond TI. Topological crystalline
insulators (TCI) are a category of materials in which their Dirac surface states are
protected by the crystalline symmetry, rather than the time reversal symmetry in TIs
[51] [52]. Dirac fermions are confirmed on the (111) surface, which is difficult to
obtain from single crystal cleavage, of the MBE grown TCIs: Pb1−x Snx Te [53] and
SnSe [54] in rock-salt structure.
23 Material Preparation and Thin Film Growth 1173

Dirac and Weyl Semimetals


The realization of topologically protected electronic states in topological insulators
triggered enormous interest to discover other manifestations of topology in solid-
state crystals, like the case of Dirac and Weyl semimetals. Unlike topological
insulators, where the Dirac fermions live only on the surface states, these particular
semimetals present linear band dispersions in their bulk bandstructure, making
topology a three-dimensional property. A Dirac semimetal is thus often referred
to as the three-dimensional analogue of graphene, whereas in the case of Weyl
semimetals, the Dirac cones appear in pairs with opposite chirality [55]. The
discovery of Weyl semimetals was particularly exciting because it was understood
as the realization of the Weyl fermion – originally a concept in high-energy
physics – in a crystalline solid. The fact that Dirac and Weyl semimetals offer a
three-dimensional topological order is particularly appealing to harness topological
features in thin film and device applications.
Most of the Dirac and Weyl semimetals reported up to date are synthesized via
solid-state chemistry methods, but remarkable efforts have been made to develop
recipes for thin film growth, in view of applications. MBE has been so far the
method of choice, as a high-quality epitaxy with low-defect concentrations is crucial
for the observation of topological features. Moreover, the ability to modify the
Fermi-energy toward the Dirac or Weyl points via doping or strain engineering is a
powerful tool that is increasingly attracting more thin film growth experts to venture
into this emerging field. Dirac semimetals grown by MBE have been reported a
few years ago in the compounds K3 Bi and Na3 Bi [56, 57] as well as Cd3 As2
[58], while Weyl semimetal thin films have been achieved very recently [13] in
the monopnictide NbP and TaP compounds. Examples of the grown structures and
its electronic properties can be found in Figs. 10 and 11.

Fig. 10 Examples of Dirac semimetals grown by MBE. (a) Crystal structure of Na3Bi, highlight-
ing the two non-equivalent Na sites in the unit cell. (b, c) RHEED patterns of the Si (111) – 7 × 7
substrate along [11-2] and of the Na3 Bi film along [110]. (d, e) Complementary RHEED patterns of
the Si (111) – 7 × 7 substrate along [110] and of the Na3 Bi film along [11-2], indicating an epitaxial
growth with an in-plane rotation of 30◦ between film and substrate. (f) Electronic bandstructure of
Na3Bi along -M, showing a clear linear dispersion characteristic of Dirac semimetals. (Taken
from [56] (With kind permission from Elsevier))
1174 A. Bedoya-Pinto et al.

Fig. 11 (continued)
23 Material Preparation and Thin Film Growth 1175

High-Temperature Superconductors
Despite more than 30 years of intensive studies, the mechanism of high-temperature
superconductors (HTC) is still a mystery. Both the copper- and iron-based super-
conductors have layered tetragonal structures, which are naturally suitable for
MBE growth. Nevertheless, since most of the HTCs contain four to five elements
(e.g., Bi2 Sr2 CaCu2 O8+δ , Ba1−x Kx Fe2 As2 ), the growth of high-quality HTC films
requires sophisticated control of beam fluxes.
The most successful MBE growth of cuprate HTCs was performed by I. Božović
and his colleagues in Brookhaven National Laboratory, where a technique called
atomic-layer-by-layer MBE (ALL-MBE) is applied [59]. ALL-MBE is developed
based on the fact that each atomic layer in a cuprate HTC only contains oxygen
and one metal element. In ALL-MBE, the real-time signals from the atomic
absorption spectroscopy (thickness measurement) are fed to the shutter controller,
which switches the molecular beams exactly at the time that one atomic layer is
finished. ALL-MBE can tune the compositions of complex cuprate HTCs freely
and can create atomically sharp interfaces between different cuprate materials [60].
Thanks to the excellent reproducibility of the growth, the scaling law between
superconducting critical temperature Tc and the 2D superfluid phase stiffness can
be revealed in a statistical way, which shows deviation from the standard Bardeen-
Cooper-Schrieffer description of superconductivity (Fig. 12c) [61].
Comparing with cuprate HTCs, the MBE growth of iron-based superconduc-
tors was realized much faster after their discoveries, such as FeSe, KFe2 Se2 ,
Sr1−x Kx Fe2 As2 , Ba1−x Kx Fe2 As2 , SmFeAs(O,F), LiFeAs, etc. [62–65]. However,
the early MBE grown films did not show better performance than their bulk
counterparts. A major breakthrough happened on the monolayer FeSe films grown
on SrTiO3 (100) substrates, in which the superconducting Tc (onset 53 K) measured
from ex situ electric transport experiments is much higher than the bulk material
(9.4 K) (Fig. 12d–g) [66]. Later, an in situ 4-probe transport experiment in UHV
environment implies a Tc higher than 100 K in as-grown monolayer FeSe films
(Fig. 12h) [67]. The mechanism of the Tc enhancement in monolayer FeSe is still
under debate and has attracted intensive research interests since its discovery [68].

Layered Transition Metal Dichalcogenides


Since the discovery of graphene in 2004 [69], 2D materials have been considered to
be candidates for future high-performance electronics beyond silicon-based devices.
Aside from graphene, the most intensively studied 2D materials are transition metal


Fig. 11 Weyl semimetals grown by MBE. (g) Crystal structure of Nb(Ta)P and schematic cross
section of MgO substrate and Nb-buffer layer used for growth. (h–k) RHEED pattern of MgO
substrate and NbP film, indicating an epitaxial growth with an in-plane rotation of 45◦ between
film and substrate. (l) Momentum-resolved photoemission spectra of NbP along A-A direction,
where the Weyl points are expected. A linear band dispersion is observed. (m) Fermi-surface of
NbP, with four electronic pockets along X and Y directions, ascribed to topological surface states.
(Taken from [13] (© ACS, with kind permission))
1176 A. Bedoya-Pinto et al.

Fig. 12 (a) Schematic diagram of a cuprate HTC sample grown by ALL-MBE [37]. (b) Scanning
transmission electron microscopy of a cuprate metal-insulator bilayer. Inset, a magnified image of
the interface (arrowed) [60]. (c) The dependence of superconducting Tc on the zero-temperature
2D superfluid phase stiffness ρ S0 in La2−x Srx CuO4 [61]. (d) STM topography and (e) atom
resolved images of monolayer FeSe film grown on SrTiO3 (100). (f) A dI/dV spectrum showing
a superconducting gap in monolayer FeSe. (g) Ex situ electric transport experiments of monolayer
FeSe capped with amorphous Si [66]. (h) In situ 4-probe transport experiments on monolayer FeSe
[67] (b, c, and h reprinted with permission from Springer Nature)

dichalcogenides (TMDC), which are a family of both semiconducting and metallic


materials with the chemical composition of MX2 (M = transition metal, X = S, Se,
or Te).
Monolayer semiconducting TMDC materials in 1H structure are promising
for the potential spintronic and valleytronic applications due to their spin-valley
coupled band structure [70]. Because of their van der Waals layered structure,
TMDC materials are usually easy to exfoliate from bulk materials. However, large-
scale fabrication relies on the thin film growth techniques, such as chemical vapor
deposition (CVD) and MBE. Most of the MBE-grown TMDC thin films use
epitaxial graphene on SiC(0001) as substrates because of its extremely low surface
energy. Observed with in situ ARPES, indirect-direct band gap transitions as the
film thicknesses decrease in MBE grown MoSe2 and WSe2 are revealed, together
23 Material Preparation and Thin Film Growth 1177

Fig. 13 (a) The lattice structure of 2H-MoSe2 . Three atomic layers form a monolayer 1H-MoSe2 .
(b) The ARPES data shows the VBM of MoSe2 changes from Γ point to K point as the thickness
decreases from bilayer to monolayer, which leads to an indirect-direct band gap transition. The
spin-orbit splitting of the bands at K point is also resolved [22]. (c) The 1D CDW at the twin
boundaries of monolayer MoSe2 [73]. (d) The lattice structure of 1T’-WTe2 . (e) The QSH band
gap of 1T’-WTe2 resolved by ARPES. The sample surface is K-doped. (f) The dI/dV spectra
measured by low-temperature STM showing enhanced density of states at an edge of 1T’-WTe2
[79] (a, b, and f reprinted with permission of Springer Nature)

with spin-orbit splittings as large as 180 meV (MoSe2 ) and 475 meV (WSe2 )
at the valence band maximums (VBM) (Fig. 13b) [22] [71]. Compared with the
CVD grown films, the MBE-grown semiconducting TMDC films have much higher
density of 180◦ twin boundaries, which are found to be metallic and hosting one-
dimensional charge density waves (CDW) (Fig. 13c) [72] [73].
On the other hand, various CDW orders and superconductivity are discovered in
the metallic monolayers of 1H-TiSe2 [74], NbSe2 [75], and TaSe2 [76]. The super-
conductivity in NbSe2 is especially interesting because of its Ising pairing, which
leads to an in-plane upper critical field much higher than the Pauli paramagnetic
limit [77].
The 2D TMDC materials also exhibit topological orders. 1T’ structured mono-
layer TMDC (Mo/W)(S/Se/Te)2 are predicted to be QSH systems, in which
spin-orbit interaction opens a band gap in the inversed bands and leads to QSH
edge states [78]. QSH band gaps are confirmed in the MBE-grown 1T’-WTe2 and
1T’-WSe2 films by ARPES experiments (Fig. 13e, f) [79] [80], while 1T’-MoTe2
turns out to be a semimetal due to its insufficient spin-orbit interaction [81].
Furthermore, two-dimensional ferroelectricity has been discovered in MBE-
grown monolayer chalcogenides very recently. Monolayer SnTe grown on graphene
is found to be ferroelectric with in-plane polarization [82]. Such 2D ferroic materials
are considered to be essential for the conception of future full-2D-material memory
devices.
1178 A. Bedoya-Pinto et al.

Comparison with Other Deposition Techniques

In this section, we will make a brief comparison between MBE and the other
thin film deposition techniques and discuss about the future development of MBE
technique.
When growing those binary compounds with group V or VI nonmetals, with
extremely clean environment, slow growth rate, and precise beam flux control,
MBE can usually achieve the best film or heterostructure quality among the
common thin film growth techniques, such as magnetron sputtering, pulsed laser
deposition (PLD), chemical vapor deposition (CVD), and electrochemical liquid
phase deposition. The chemical stoichiometry can be almost perfect in these MBE
grown films, thanks to the three-temperature method mentioned above. The growth
of pseudo-binary compounds, in which several metal or nonmetal elements can
freely substitute each other in the crystal lattice, such as (Al/In/Ga)(N/P/As),
also follows this principle. With specially designed effusion cells, growing the
compounds with group IV or VII nonmetals is also possible. In ALL-MBE, despite
the complicated chemical composition of cuprates, the three-temperature method is
still valid when growing each atomic layer (one metal element plus oxygen).
However, when the desired material is not in a layered structure and possesses
multiple metal elements, such as Heusler compounds (chemical composition XYZ
or X2 YZ, in which two or three metal elements are involved), it becomes difficult
to match the beam fluxes. Unlike sputtering or PLD, which typically uses a single
target to get all the metal elements, it is often impossible to obtain stoichiometric
metal element fluxes from a single effusion cell in MBE, because in this case, the
whole source material in the crucible is in a thermal equilibrium, and the vapor
pressures for each metal element are different at a certain temperature. Depositing
compound films from a single effusion cell is only possible when the source
material evaporates as stoichiometric compound molecules, such as NaCl, MgO,
Pb1−x Snx Te, etc. On the other hand, a target under the local bombardment/heating
in sputtering or PLD is far away from thermal equilibrium; thus, the elements at the
sputtered/heated spot can be evaporated altogether. It should be noted that although
the growth of Heusler compounds is relatively difficult for MBE, there are indeed
considerable amount of reports on MBE-grown Heusler compound thin films, whose
magnetic properties are close to the bulk materials [83] [84].
Comparing with the other thin film growth techniques, another drawback of
MBE is its relatively low flexibility of changing the source materials. Since UHV
environment can only be achieved after a bake-out process, it usually takes several
days to change source materials in MBE systems.
From the scientific research’ point of view, two future prospects of MBE
can be immediately recognized as important: (i) As next-generation electronics
is pursuing 3D structures, heterostructures beyond stacked thin films become
important recently, for example, lateral p-n junctions at the boundaries between
monolayer WSe2 and MoSe2 [85], and (ii) all in situ growth and characterization
in UHV environment have been proven to be a powerful approach in developing
novel functional materials. This includes not only surface science techniques (STM,
23 Material Preparation and Thin Film Growth 1179

ARPES, XPS, etc.) but also electric transports (local 4-probe), optical measurements
(photoluminescence, Raman spectroscopy, microwave impedance imaging, etc.),
and magnetic measurements (two-coil mutual inductance).

Pulsed Laser Deposition

Introduction

Pulsed laser deposition (PLD) is one of the most used techniques for depositing thin
films. It is a physical vapor deposition technique which is ideally suited to deposit
thin films of complex or multicomponent oxides onto a substrate as this technique
congruently transfers the composition of the target to the film. PLD technique gained
enormous popularity following the discovery of high-temperature superconductivity
in perovskites [86] when it was demonstrated that high-quality perovskite thin films
with bulk-like properties could be deposited with PLD [87]. During PLD process,
the target material is vaporized/ablated with a high-intensity excimer radiation pulse
whose width is of the order of tens of nanoseconds and whose energy is of the order
of joule. As a result, a jet of particles, commonly denoted “plume” due to its form
(see Fig. 15), is ejected from the target surface and expands away from the target
with a strong forward-directed velocity distribution. The ablated particles condense
on a substrate usually placed opposite to the target; this process takes place in a
vacuum chamber – either in the high vacuum range or in the presence of some
background gas. The laser pulses are guided to the vacuum chamber by means of
optical devices, such as mirrors and lenses, which in addition focus the beam to the
target, optimizing the energy density of the laser pulses. While the laser pulses are
hitting on its surface, the target is usually rotated with a constant speed to achieve
a homogeneous ablation process. The possibility of a multi-target rotating wheel
in the vacuum chamber enables more efficient and complex processes: multilayers
and alloy films can be grown from elementary targets by moving them alternately
into the laser focal point. The high energy density used in a typical PLD process
is able to ablate almost every material, and by controlling the process parameters,
high-quality films can be grown reliably at a much faster rate compared to other
growth techniques such as MBE. Another known advantage of the PLD technique
is the stoichiometric material transfer from target to film, such that compounds with
highly complex compositions can be grown using only one target source.
Although the pulsed laser deposition process is conceptually simple, controlling
the dynamics of the film growth is not necessarily a straightforward task, because
of the large number of interacting parameters that govern the growth process and
hence the film properties, such as:

• The laser parameters (working wavelength, pulse energy, width, and repetition
rate)
• The substrate temperature and rotation speed
• The structural and chemical composition of the target material
1180 A. Bedoya-Pinto et al.

• The chamber pressure and the chemical composition of the buffer gas
• The geometry of the experiment (incident angle of the laser, incident angle of the
plume, distance between target and substrate)

Being able to control the laser and deposition parameters for a given system is the
clue to profit the advantages of the PLD technique. In practice, parameters such as
the experiment geometry are usually optimized for a given system and kept constant,
while other parameters such as laser flux, substrate temperature, chamber pressure,
and background gas are systematically varied in order to investigate their influence
on the film growth. The PLD chamber base pressure is typically in high vacuum
range. The PLD is not well suited to grow films from highly thermally conducting
or metallic targets as the incident laser energy into the target is quickly dissipated
leading to a low deposition rate. A schematic view of a typical configuration of a
PLD apparatus is sketched in Fig. 14.

Fig. 14 Schematic view of a typical PLD apparatus


23 Material Preparation and Thin Film Growth 1181

Target Preparation and Manipulation

The two important criteria which determine the quality of the PLD targets are their
density and stoichiometry. It is important that the density of targets is close to or
better yet exceed 90% of the theoretical value. Such dense targets are fabricated by
hot or cold pressing/sintering of powders of material with requested stoichiometry.
During the PLD process, the elemental ratio of the target constituents except the
oxygen content is transferred to the substrate, and hence it is essential that target
has the appropriate stoichiometry. The ablation of the target is localized to the laser
spot incident on it, and the target utilization is improved by rastering the laser spot
on the target surface or by target rotation or by combination of both laser rastering
and target rotation. The PLD targets are routinely resurfaced by polishing away
surface asperities created during extensive target use. Dry polishing with a fine-grit
sandpaper or wet polishing with an alcohol-based polishing mixture is preferred.
Aqueous polishing mixtures are avoided to minimize any risk of any water retention
within the targets, especially so for porous targets. The metallic targets for PLD are
fabricated by similar processes to those used for sputtering targets and are described
in the sputter deposition section.

Laser

Excimer lasers are most widely used to ablate material during PLD process as they
provide laser beam at higher power, higher repetition rate, and superior flat-top-
beam profile. Excimer is an acronym for EXCIted diMER. diMER is a misnomer
in some cases where KrF is used as a halogen gas but is used because of historical
reasons (emission in first excimer lasers was from dimers of Xe). Also note that
KrF is not a stable molecule in its ground state and is only formed in the excited
state within high-energy electrical discharge in the laser cavity. The excimer lasers
used halogen gas-filled cavity and laser with wavelength of 157 or 193 or 248 or
308 or 351 nm depending on the halogen gas used, F2 or ArF or KrCl or KrF or
XeCl or XeF, respectively. These lasers do require expensive gas handling system
due to health concerns with the use of halogen gas. The most common wavelength is
248 nm generated using KrF. The shorter-wavelength photons have higher energy, so
they have higher ablation efficiency and thus result in a higher deposition rate. The
pulse width of 25 ns is typical for an excimer laser, and they can deliver power of
over 1 J per pulse. An alternate to the excimer laser is the solid-state Nd:YAG laser
as it is relatively inexpensive and easier to maintain. The fundamental wavelength
of the Nd:YAG laser is 1064 nm, and hence its optical frequency can be doubled,
tripled, quadrupled, or quintupled. Though, the most common setup with this laser
is to use a wavelength of 266 nm (frequency quadrupled) or 355 nm (frequency
1182 A. Bedoya-Pinto et al.

tripled) with this laser. Excimer lasers (KrF, ArF, XeCl) are widely used to deposit
oxide films because of the larger absorption coefficient and small reflectivity of
materials at their operating wavelengths.

Ablation Process

The energy density within the focused laser beam spot is extremely high (range
of 0.1–2 J/cm2 ) and which due to the short pulse duration (∼10 ns) corresponds
to instantaneous power densities in the range of 107 –108 W/cm2 . The high power
density can locally heat the target material to temperatures of even few thousand
Kelvin causing thermal evaporation of the target material in the initial stages [88].
During laser ablation process, a dense cloud of vaporized material is formed which
contains a mixture of ions, molecules, atoms, and even small particulate in extreme
cases. This material cloud continues to absorb energy from the laser beam and
consequently undergoes a sudden expansion into a background gas or vacuum to
form the laser plume (Fig. 15). This laser plume expands away from the target
and travels toward the substrate surface leading to the eventual deposition of a thin
film of the target material. The kinetic energy of the plume species from metallic
targets ranges between a few to ∼100 eV in vacuum [89]. Use of background
gas (pressures greater than mTorr) reduces the kinetic energy of the plume species
and also increases the number of chemical reactions between plume species and
the gas molecules. The mean kinetic energy of the species being deposited in
background gas is of the order of 10 eV which is comparable to that during
sputter deposition. Some of the lighter elements like oxygen or nitrogen tend to
have higher kinetic energy and different angular distribution [90], and hence it is
customary to supply these elements via the background gas so as to deposit film of
the desired composition. The background gas broadens the plume, and the extent
of this broadening depends on the gas pressure. If it is essential to maintain a low
background gas pressure, then it is possible to obtain thin of desired composition by
adjusting the target composition.

Fig. 15 Photographs of different plasma plume shapes generated in a laser ablation process. The
size, shape, and color depend on the laser parameters and the target material. Examples of widely
used material targets such as carbon (middle panel) and ZnO (right panel) are highlighted
23 Material Preparation and Thin Film Growth 1183

Optics

The laser beam is typically focused and incident on the target at angle ranging from
30 to 45◦ from its surface. The laser is focused using a focusing length needed
to focus the laser beam on to the target. In some PLD systems, raster mirror is
added in the beam path downstream of the focusing lens to deflect the partially
focused beam on the target. The raster mirror is motorized to change the deflection
angle during deposition such that the beam spot scribes a line on the target surface.
This line is offset from the center of the target which is continuously rotated
during deposition. This laser beam rastering improves chemical compositional and
thickness uniformity and target usage. During rastering of the laser beam, the laser
beam incidence angle on the target varies which affects the laser beam energy
density on the target consequently causing a variation in ablation. This disadvantage
can be partially overcome by substrate rotation during deposition. Though laser
output energy of the incident laser beam on the target can be controlled by changing
the high voltage used during the emission process, this is not preferred as it also
changes the beam profile which in turn affects the incident laser intensity on the
target. It is better to operate the laser at constant energy and reduce the beam energy
using an attenuator mounted external to the laser. The attenuator has minimal impact
on the laser beam profile, and consequently a better control of the laser ablation
is achieved. Sometimes optical slits are used to remove inhomogeneous parts of
the laser beam. All the optical components are coated with anti-reflection coating
appropriate for the wavelength of the laser to improve light transmitivity by virtually
eliminating unwanted reflections.

Thickness Monitoring

High-pressure reflection high-energy electron diffraction (RHEED) is used as in


situ monitor of the deposited film thickness and crystallinity. A collimated and
small-size (typically <1 mm) electron beam is incident on the surface at grazing
angle typically within the range of 1–3◦ . The specularly reflected beam along with
the diffraction spots is imaged on a phosphor screen placed on the opposite of
the source gun using a low-noise CCD camera. The observed diffraction pattern
is analogous to the X-ray diffraction pattern and provides information on crystal
structure of the film and/or substrate. Thus signal from amorphous, polycrystalline,
and single-crystalline or highly textured film is diffuse background, ring-lie pattern,
and a first-order Laue circle of spots, respectively. The specular spot intensity
oscillates if the film thickness increases layer-by-layer (Frank-van der Merwe
growth). A single RHEED oscillation corresponds to growth of a single layer of
the film, and the minimum in the oscillation corresponds to formation of half the
layer. The electron guns use a tungsten filament to generate electron beam with
energies between 20 and 60 keV. The stable operation of the filament limits the
pressure in the region surrounding the filament to <∼10−5 torr. The background
1184 A. Bedoya-Pinto et al.

gas pressure during the PLD process tends to be several orders of magnitude higher
(ranging from 1 to 500 mTorr) which necessitates the use of a two-stage differential
pumping to maintain high vacuum in proximity of the tungsten filament. Such
differentially pumped RHEED is called high-pressure RHEED. The scattering of
the electron beam by background gas is minimized by locating the exit aperture
of the high-pressure RHEED and also the phosphor screen close to the substrate
(roughly within few cm). Moreover, a shutter is used to protect the phosphor screen
when the RHEED is not in use. A quartz crystal microbalance (QCM) can also be
used to monitor the growth rate of the PLD process. These monitors can be used
prior and post deposition to measure the growth rate.

Substrate Heating

Substrate temperature prior to growth (i.e., substrate conditioning), during depo-


sition, and post deposition is an important factor which determines the quality of
the PLD thin films. During deposition, substrate temperature determines surface
mobility and re-evaporation of the deposited material. The substrates are heated in
background gas of oxygen to temperatures as high as 1000 ◦ C, and this determines
the material choices for wafer holders and substrate heaters. SiC is stable to these
high temperatures in oxidizing environment and can be mounted in close proximity
of the substrate to radiatively heat it. A potential heater design includes a serpentine
SiC heater element surrounded by multiple heat shields to localize the generated
heat. Alternatively, high-power diode laser (∼500 W) with its radiation incident on
the back side of the substrate can be used to heat the substrates. Such laser heating
is advantageous as the heater components are outside the vacuum chamber which
makes the maintenance easy. Moreover, the heater radiation is well collimated and
localized to the substrate so that there is no inadvertent outgassing of the vacuum
chamber components.

Selected Applications

Growth of VO2 using PLD. VO2 is an archetypal correlated oxide with a metal
to insulator transition (MIT) near room temperature, and hence it has been exten-
sively explored to develop novel electronic devices. Excellent quality epitaxial,
single-crystalline, and stoichiometric thin films of VO2 were deposited from
polycrystalline VO2 or V2 O3 targets by pulsed laser deposition (PLD) technique
on various substrates [91]. This study used a laser energy density of 1.3 J/cm2 ,
a repetition rate of 2 Hz, a target to substrate distance of ∼7.1 cm, and oxygen
deposition pressures of at least 9 mTorr. Film quality and properties were not much
affected for oxygen pressures that were varied between 9 and 15 mTorr. Growth
temperatures were substrate-dependent, and the largest change in resistance of three
to four orders of magnitude at the metal to insulator transition (MIT) was found
23 Material Preparation and Thin Film Growth 1185

for deposition temperature of 400 ◦ C, 500 ◦ C, and ∼700 ◦ C for TiO2 (001) and
(101), TiO2 (100) and (110), and Al2 O3 (0001) and (10-10) substrates, respectively
(see Fig. 16a). The dependence of MIT temperature on various substrates is a direct
consequence of a different strain state in the VO2 films. Moreover, AFM studies
indicated that with optimized growth conditions, the deposited films were smooth
and without any particulate debris typically seen in PLD films. High-resolution
XRD peaks of VO2 along with the Kiessig fringes obtained at Stanford Synchrotron
Radiation Laboratory (Fig. 16 and Ref. 92) illustrate high quality of the epitaxial
growth on TiO2 (001) and TiO2 (110) substrates. Substrate preparation also plays
an important role in determining the properties of thin VO2 films grown on such
single-crystalline substrates [93].
Growth of perovskites using PLD. High-temperature superconductors with criti-
cal temperatures higher than liquid nitrogen temperature (77 K) are based on hole-
doped cuprates having perovskite structure containing copper oxide planes [86].
PLD is the technique of choice to deposit thin films of high-temperature super-
conductors. Perovskite compounds, ABO3 , contain alternating layers of AO and
BO2 planes along its c-axis. These two planes are chemically distinct, and hence
the surface properties of perovskites are termination-dependent. Recently, high-
mobility electron gas was observed at the interface of nominally insulating LaAlO3
and SrTiO3 when the substrate was TiO2 terminated [94]. This interface was
prepared by PLD growth of LaAlO3 on SrTiO3 single crystal substrates. The
LaAlO3 films were deposited using a KrF excimer laser with fluence of 1 Jcm−2
lasing as repetition rate of 2 Hz, while the SrTiO3 substrate was at 800 ◦ C. This
study illustrates that high-quality perovskite films with precise thickness control
with single-unit cell resolution can be grown using PLD.

Fig. 16 (a) Metal-insulator transition of VO2 thin films grown by PLD on different oxide
substrates. (b) X-ray diffraction pattern of a VO2 film grown on TiO2 (001), indicating epitaxial
growth with a high crystalline quality, as evidenced by the appearance of Kiessig fringes along the
(002) reflection. (Panel (b) is taken from [92])
1186 A. Bedoya-Pinto et al.

Magnetron Sputtering Deposition

Introduction

Sputter deposition is the most widely used thin film deposition technique as it is
capable of depositing very-high-quality films in thickness range from angstroms
to microns even on large substrates at a highly competitive cost. This technique
uses energetic gas molecules (atoms and/or ions) to dislodge atoms from a target
surface which then travel through several inches of sputter gas onto a substrate
surface. The sputtering sources are operated in high vacuum or ultrahigh vacuum
chambers depending on the process demands. Sputtering can be classified into two
primary categories: magnetron sputtering and ion beam sputtering. The discussion
here excludes DC diode and triode sputtering as their disadvantages such as
low deposition rate, high gas densities, high discharge voltages, etc. make them
impractical in modern deposition tools. The planar magnetron which included a
magnet assembly underneath a target cathode was first introduced by J. S. Chapin
in 1974 [95], and since then his original design has been significantly upgraded for
improved performance [96, 97]. The sputtering process uses an inert gas during
deposition. Argon gas is the preferred choice due to its ample availability in
ultrahigh purity (99.999%) at an affordable price and also as its atomic weight is
not too low compared to most target materials. The targets are in solid form and
conform to the magnetron source size and shape.

Material Targets

The pure metal or metallic alloy targets are fabricated by arc melting/casting
high-purity elemental metals in an inert Argon atmosphere as this method yields
targets with densities close to the theoretical values and with minimum amount
of oxygen inclusion. The metal targets can also be fabricated by sintering and
vacuum induction melting. The dielectric targets are fabricated by hot or cold
pressing/sintering of powder of dielectric materials. It is important that density of
the dielectric targets be close to or better even exceed 90% of the theoretical value.
The target fabrication method is known to influence the properties of deposited
thin films, and hence it is necessary to empirically determine the preferred target
fabrication method. During sputtering, the targets are typically water cooled to
maintain stable operation, and hence in many cases, the targets are backed with
a copper backing plate. The adhesion between the target and backing plate can be
achieved with either indium or epoxy bonding. The indium bonding is preferred as
indium is malleable which prevents its cracking. The low melting point of indium
will lead to failure of the bond if the targets get too hot during operation or bake-
out. Epoxy or elastomer bonding is used where indium bonding is not possible as
the target material is low melting or is temperature sensitive. Furthermore, the target
bonding to a backing plate also helps maintain integrity of brittle targets.
23 Material Preparation and Thin Film Growth 1187

Magnetron Sources

The magnetron sputter gun or cathode houses a series of permanent magnets whose
field lines localize the process plasma around the target surface by providing
drift paths for the electrons emanating from the target. The actual arrangement
of the permanent magnets within the sputter gun not surprisingly depends on
the gun manufacturers. The need to maintain sufficient magnetic field above the
target surface limits the target thickness particularly for target made from magnetic
materials to about half to a third of the thickness used for a non-magnetic target.
The target does not wear uniformly with use, but instead a trench or a groove is
formed in the region where most energized Argon ions impinge on the surface. In
the case of circular magnetron source, the groove is also circular and is referred to
as a “racetrack.” This groove deepens with target usage and eventually cuts through
the entire target thickness. The target utilization is low and less than 50% in the best
case. Hence, the precious metal targets are refurbished by adding fresh material to
an exhausted target. Interestingly, for difficult-to-light magnetic targets, mechanical
modification of the target to include a shallow groove promotes stabilization of the
process plasma when the target is in its pristine state. The target is clamped to the
sputter gun though a magnetic target could be just placed atop the gun without a
clamp. A negative voltage of several hundred volts is applied to the target surface
during the sputtering process. The target voltage decreases as the “racetrack” groove
deepens, and hence target wear/life can be monitored by careful tracking of this
voltage. It is essential that the target be electrically isolated from the ground/anode
surface/shield which surrounds the target. The spacing between these shields and the
gun is termed as the “dark space” gap and is around 1 mm depending on the type of
the magnetron source. This gap is adjusted to match the stable operating sputter gas
pressure. Too small a gap causes shorting, and if it is too large, then plasma is formed
within this region leading to sputtering of the clamping ring material which can
contaminate the intended thin film. Figure 17 shows a computer-generated picture
of a magnetron source and the associated components which are described in detail
below.

Magnetron Operation

During stable operation, the electrons spiraling along the gun’s magnetic field lines
ionize the sputter gas to positive ions forming a magnetron plasma confined close
to the target. This plasma emits a glow of light. The negative voltage applied to
the target attracts the positive ions which impact the surface at sufficient energies
creating a collision cascade at the target surface. As a result, photons, secondary
electrons, and target atoms/molecules are ejected from the surface. These ejected
atoms/molecules along their way toward the substrate undergo numerous collisions
with the sputter gas atoms which deflect them and result in loss of their kinetic
energy. Some of them are even backscattered to the target surface. This deflection
1188 A. Bedoya-Pinto et al.

Fig. 17 Computer-generated picture from SolidWorks 3D model of a magnetron deposition


system. Only the hardware components relevant for deposition with the magnetron source are
included. (Courtesy: Chris Lada)

of the ejected atoms/molecules decreases as the atomic weight of the sputtered


atoms increases or if the sputter gas pressure is reduced. This effect can be used
advantageously to fine-tune the composition of a sputter deposited film from a
single alloy target (see Fig. 18). Moreover, the concentration of the heavier element
in the sputtered film is higher than its content in the target. In spite of these
deflections, the sputtering is predominantly a line-of-sight deposition method. The
number of collisions between ejected atoms/molecules and the sputter gas increases
as the target-substrate distance is increased. At larger separation, the deposition
rate goes down by inverse square relationship, whereas the film thickness and
composition uniformity improve. The optimum target-substrate distance depends
on the magnetron source geometry. For circular sources, the plasma is contained
within the target radius, so that the recommended distance should be greater
than the target diameter. At very short separation distances, interaction between
magnetron plasma may cause substrate heating, arcing, and damage. The use of
a magnet configuration within the magnetron with net magnetic field toward it
results in ion bombardment of the substrate during film growth which leads to
the deposited films being denser and with preferential texture. In the opposite
case where the net magnetic field is directed away, ion bombardment is lowered
on the substrate. Such configured magnetrons are referred to as “unbalanced”
magnetrons [98].
23 Material Preparation and Thin Film Growth 1189

Fig. 18 (a). Pt content as determined by Rutherford backscattering (RBS) for a series of dc


magnetron sputtered deposited thin films of PtMn alloy as a function of Ar sputter gas pressure.
Two different composition PtMn targets were used. All films have higher Pt content than the target,
and this increases as the sputter pressure is increased. (b). Thickness of the PtMn films as function
of the Ar sputter gas pressure. The deposition rate is higher at lower sputter gas pressure

Chimneys and Shutters

Chimneys are used to provide shielding to contain sputtered material preventing


their deposition on unwanted chamber surfaces and to locally enhance the gas
pressure near the target surface via gas injection into the chimney. Thickness of the
deposited film can be precisely controlled by a shutter operated with a pneumatic or
motorized actuator. High actuator speed for the shutter is desirable to allow precise
control of thin layers which could be ∼1 monolayer thick. Since the chimneys and
shutters are constantly exposed to source plasma, there is significant material build
up on them. Some materials are prone to formation of flakes. The bane of magnetron
1190 A. Bedoya-Pinto et al.

sputter deposition is arcing or unstable source operation caused by these material


flake shorting the source to the ground shields. Though tapping the source with a
rubber mallet may give temporary relief, it is necessary to physically remove the
flake. The chimneys and shutters have to be thoroughly cleaned on a regular basis,
and sand blasting is an effective method for removal of material deposits from them.

Reactive Magnetron Sputtering

Oxide or nitride thin films can be deposited from metal targets by addition of
oxygen or nitrogen, respectively, to the inert Argon sputter gas. This method is
referred to as “reactive” sputtering [99] and is a cheaper and more convenient
method than RF sputter to deposition of oxide or nitride thin films from dielectric
targets. During such process, the reactivity of the oxygen or nitrogen molecules
within the plasma, contained above the target surface, is enhanced and their
reaction is promoted with the target surface, forming an oxide or nitride layer
which then is sputtered away. A caveat, the dielectric layer formed during reactive
sputtering, lowers the sputtering rate with time as the surface becomes more and
more insulating and eventually leads to frequent arcing between the target and
ground shield. Pulsed DC power supplies which generate positive voltage spikes
at set frequency can be used to minimize buildup of dielectric layer on the target
surface. These power supplies also include arc suppression capability to detect and
extinguish an arc. When an abrupt fall in a target voltage of ∼50–150 V which is
associated with an arc formation is detected, the power applies a reverse voltage
pulse to quench the arc. The metal target surface can be regenerated by sputtering
the surface with pure inert gas. The composition of the oxide or nitride film (i.e.,
its oxygen or nitrogen content) can be controlled by varying the reactive sputter gas
mixture (see Fig. 19).

RF Magnetron Sputtering

An alternate way to deposit thin dielectric films is the use of RF (radio-frequency)


magnetron sputtering from a dielectric target. Here, the alternating electric potential
avoids buildup of electric charge at the dielectric target surface. The RF process does
not rely on trapping of secondary electrons above the target to sustain the plasma.
Instead, the electrons are accelerated and decelerated over a sufficient distance by
the RF to sustain the plasma. In addition, the interaction between the electrons and
the source magnets is weak, so the RF plasma is not tightly confined which results
in a less pronounced target erosion pattern (racetrack formation) than in the DC
case. The RF sputtering plasma is stable at lower pressures as low as 1 mTorr.
This technique necessitates the use of RF power supply (typical source frequency of
13.56 MHz), matching network module mounted close to the magnetron source, and
a tuner control module, and all of which are expensive. The high-reflected power
or RF operation without a plasma can heat the source resulting in damage to the
23 Material Preparation and Thin Film Growth 1191

Fig. 19 The composition of Mn nitride films deposited with reactive magnetron sputtering
using a mixture of Ar and N2 as sputter gas from a pure Manganese target. The composition
was determined by Rutherford backscattering (RBS). To accurately determine the composition,
200 Å films of Mn nitride were deposited on graphite substrate at room temperature and capped
with 100 Å Pt to prevent their oxidation upon exposure to ambient. The graphite substrates
which contain predominantly carbon, an element with lower atomic number than nitrogen, allow
determination of nitrogen content of the Mn nitride films to within +/−0.5%

source itself. During RF sputtering, the relative duty cycle is lower than with DC
sputtering which results in a lower deposition rate. The targets of dielectric materials
are always bonded to the Cu backing plate to facilitate removal of heat generated at
the target surface during operation.

Growth of Alloy Films

The deposition of alloy thin films is achieved by using an alloy target from a single
magnetron source. Due to differences in sputter yields, transmitivity through sputter
gas, and the chamber geometry, the composition of the sputter deposited alloy film
never matches that of the target. The composition of the alloy thin films can be varied
by changing sputter gas pressure or target-substrate separation (mentioned above)
or in addition by changing the off-axis angle (i.e., the angle between substrate and
target normal). An alternate means for deposition of alloy thin films is the use of
multiple sputter sources arranged in a circular pattern aimed at a common focal
point, so-called “confocal” geometry (see Fig. 20). The compositional tuning and
also growth of multilayers can be easily achieved in this geometry. The uniformity
of composition and thickness of the thin films to better than +/−1% across the entire
1192 A. Bedoya-Pinto et al.

Fig. 20 Illustration of a confocal geometry with three magnetron sources and all the associated
components as installed in one of the deposition systems at the IBM Almaden Research Center.
Each magnetron source has its own individual chimney and shutter to prevent cross contamination.
A separate shutter is used to control film thickness. (Courtesy Chris Lada)

substrate can readily be achieved by rotating the substrate during thin film growth.
Sputter deposition rate can be monitored by a quartz crystal monitor (QCM) prior
and post growth or even during deposition by appropriate placement of the QCM
within the deposition chamber. Heating or even cooling of substrate to temperatures
as high as 1000 ◦ C or as low as −195 ◦ C is used to achieved desired thin film texture,
crystallinity, and smoothness. At typical magnetron sputter conditions, the ejected
target atoms arrive at substrate with kinetic energies of a few eV which is an order
of magnitude higher than in molecular beam epitaxy.

Ion Beam Sputter Deposition

In ion beam sputtering, the target atom kinetic energy upon arrival to the substrate is
higher than in magnetron sputtering as the sputter gas pressure within the deposition
23 Material Preparation and Thin Film Growth 1193

chamber is almost an order of magnitude lower. Thus, in ion beam sputtering,


the mean free path between collisions for ejected atoms/molecules is typically
longer than the target-substrate separation. Consequently, the ion beam sputter
deposited films are denser. In ion beam deposition (IBD), a stream of inert gas
atoms are extracted from a plasma source and directed onto a target surface at
energies between ∼50 and ∼1000 eV. Argon is the most commonly used inert gas
during this process. The energetic Ar atoms sputter the target sputter with ejected
atoms/molecules being deposited on the substrate.

Ion Beam Deposition Source

Typical ion beam sputter source consists of a discharge chamber where plasma is
generated by radio-frequency (RF) or by direct-current (DC) discharge [100]. In
the RF source, an inductor coil surrounds the discharge chamber which is made
of dielectric material. In the DC source, the discharge chamber contains a hot
filament for electron emission as a cathode and a surrounding anode. The DC source
is cheaper but needs more maintenance as the source filament has a limited life.
The ions from the discharge chamber are extracted with either a two- or three-
grid assembly. (The discussion here ignores the grid-less sources as they lack ion
beam trajectory control and are not used for ion beam sputter deposition.) The
grid closest to the discharge chamber is referred to as the beam grid, and the
voltage (Vb ) applied to it is positive and equals the beam energy. The next grid
called the accelerator is at negative voltage (Va ) and accelerates the extracted ions.
The third grid in the three-grid assembly is the decelerator grid and is at ground
potential. The decelerator grid reduces beam divergence by almost a factor of
two, suppresses electron back streaming, and minimizes deposition of sputtered
ion within the source. Not surprisingly, the three-grid assembly is more expensive
than the two-grid assembly. The extracted ions are next bombarded with electrons
from a neutralizer to neutralize them, and this also reduces the beam divergence by
suppressing the mutual repulsion between the ions. Most grid assemblies are made
from molybdenum or graphite. The molybdenum grids are structurally more stable
and can be dished to shape the beam (i.e., focusing or defocusing optics). These
grids are used in thin film deposition application. The graphite grids are flat and
structurally fragile and are used in application where a collimated beam is desired.
The beam current (Ib ) measured at the beam grid is related to the etch or sputter
rate of the target. The beam current at fixed beam energy (Vb ) can be increased by
increasing the accelerator voltage (Va ). The sputter/etch rate is material-dependent
and shows a weak relationship to the melting temperature of the target material.

Ion Beam Deposition Geometry

The IBD geometry requires several considerations. The grid assembly is designed
such that almost 99% of ion beam extracted from the source is incident on the target.
1194 A. Bedoya-Pinto et al.

Fig. 21 Computer-generated picture illustrating the ion beam deposition geometry from a RF
source as mounted along with the associated components in a deposition system at the IBM
Almaden Research Center. The target manipulator houses five targets each 3 inches in diameter.
(Courtesy Chris Lada)

The incident angle of this beam on the target maximizes the sputter yield and is
usually between 40 and 50◦ from target surface. The substrate is located to close to
or along the target normal as the maximum amount of sputtered material is ejected
along this direction. There is a forward lobe along which contains atoms sputtered
at higher energies and intensity, and it is desirable that this lobe is not incident on
the substrate. The shutter used to control thickness of the deposited thin film has to
be placed such that no incident ion beam impinges on top of it as this will sputter
the shutter material on to the substrate. The neutralizer is mounted to the side of
the IBD source with its axis orthogonal to the source axis. Figure 21 shows the
IBD geometry which is used for ion beam sputter deposition in a sputter deposition
system at the IBM Almaden Research Center.

Ion Beam Deposition Targets

The IBD target are usually circular and tend to be significantly larger than the
source size due to geometric considerations and the inevitable beam divergence. The
targets are clamped to the target holder by grabbing them along their edges to avoid
sputtering of clamping material which will contaminate the deposited thin films. The
targets are actively cooled to conduct away the energy dumped into the target during
the sputtering process. During the ion beam sputtering, neutral atoms are directed
23 Material Preparation and Thin Film Growth 1195

to the target surface, and hence this process is equally effective for deposition of
dielectric films. The targets may be bonded to the copper backing plate and usually
are for dielectric or brittle materials. The IBD targets wear uniformly with use, and
life of the target is significantly longer than in magnetron sputtering. This is highly
advantageous for deposition of precious materials.

Ion Beam Deposition Source Operation

The source parameters which are controlled during operation are beam voltage,
accelerator voltage, and the sputter gas. The sputter/etch rate of a target is given by
Child’s law [101] which relates the beam current to the total voltage (Vb + Va )3/2 .
Thus, for example, at a set beam energy, the etch/sputter rate can be increased by
raising the accelerator voltage. Such voltage variation allows change in etch/sputter
rate by over two orders of magnitude. The beam divergence is related to the Va /Vb
ratio, and the beam divergence almost doubles as the ratio increases from 0.1 to
1.0 [100]. The operation of the source at high ion currents causes wear to the
accelerator grid as part of the extracted ion beam collides with this grid instead of
passing through it. The alignment of the ion optics (i.e., grid apertures) is important
to minimize grid wear due to direct ion impingement. The etch/sputter rate is only
weakly influenced by the type of inert gas used during the process (i.e., argon versus
krypton or xenon). Reactive IBD is only possible with the RF source and is achieved
by introducing a premixed inert plus either oxygen or nitrogen gas into the discharge
chamber. The ratio of inert gas to oxygen/nitrogen gas is adjusted to deposit
oxide/nitride films of desired composition. Tuning of film composition during IBD
from an alloy target is also possible by changing target-substrate distance and by
changing the off-axis deposition angle.

Selected Applications

Sputter deposition has been the method of choice for deposition of thin films
in numerous industrial branches such as magnetic storage, semiconductor, opto-
electronics, automotive, and solar energy. Several generations of hard disk drive
products have entirely relied on magnetron sputtering to fabricate the read head
sensor which is a key component of the drive. This was due to the fact that the
giant magnetoresistance (GMR) of magnetic multilayers deposited by magnetron
sputtering is among the highest recorded GMR values comparable even to films
deposited by MBE [102]. The magnetic tunnel junction (MTJ) with the MgO tunnel
barrier which shows giant tunnel magnetoresistance (TMR) of several hundred
percent can be grown by reactive magnetron sputtering [103], and this TMR is
comparable to the single-crystalline MgO tunnel barrier grown with MBE [104].
Of late, the highest TMR with MgO tunnel barrier used in magnetic random
access memories (MRAM) is obtained by growing this layer using RF sputtering.
In magnetic storage industry, the read head has evolved from magnetoresistive in
the early 1990s to GMR head in 1997 to MTJ head in 2008. This has led to an
1196 A. Bedoya-Pinto et al.

Fig. 22 Areal density perspective for HDD/Flash memory technology. The relatively low produc-
tion costs related to sputtering processes of GMR/TMR devices have been essential to introduce
such products in the market while triggering a boost in magnetic storage capacity. (Courtesy: Ed
Grochowski [105])

exponential increase in the magnetic storage areal density per inch square, as shown
in Fig. 22, where HDD and Flash technology are compared. This is perhaps the most
technologically relevant impact of the sputtering deposition technique. Its affordable
cost compared to other deposition methods has triggered the successful introduction
of magnetic storage units in the industrial market.

References
1. Tsao, J.Y.: Materials Fundamentals of Molecular Beam Epitaxy, pp. 13–41. Academic Press,
San Diego (1993)
2. Binh, V.T.: Surface Mobilities on Solid Materials, Fundamental Concepts and Applications
NATO ASI Series, Series B, Physics, vol. 86. Plenum Press, New York (1981)
3. Lagally, M.G.: Kinetics of Ordering and Growth at Surfaces NATO ASI Series, Series B,
Physics, vol. 239. Springer US, Boston (1990)
4. Mai, D.D.: Ferromagnetismus bei Raumtemperatur in mehrphasigen (Ga,Mn)N Schichten
und Heterostrukturen. p. 20, Dissertation, Georg-August Universität Göttingen (2009).
https://ediss.uni-goettingen.de/handle/11858/00-1735-0000-0006-B499-0 under CC-BY-ND
https://creativecommons.org/licenses/by-nd/2.0/de/
5. Stoica, T., Meijers, R., Calarco, R., Richter, T., Lüth, H.: MBE growth optimization of InN
nanowires. J. Cryst. Growth. 290, 241–247 (2006)
6. Calarco, R., et al.: Nucleation and growth of GaN nanowires on Si(111) performed by
molecular beam epitaxy. Nano Lett. 7, 2248–2251 (2007)
23 Material Preparation and Thin Film Growth 1197

7. Martin, P.M.: Handbook of Deposition Technologies for Films and Coatings, Science,
Applications and Technology, 3rd edn, pp. 93–134. Elsevier, Oxford, UK (2010)
8. Karpiński, J., Porowski, S.: High pressure thermodynamics of GaN. J. Cryst. Growth. 66(1),
11–20 (1984)
9. Miskys, C.R., Kelly, M.K., Ambacher, O., Stutzmann, M.: Freestanding GaN-substrates and
devices. Phys. Status Solidi C. 0(6), 1627–1650 (2003)
10. Gooth, J., et al.: Thermal and electrical signatures of a hydrodynamic electron fluid in tungsten
diphosphide. Nat. Commun. 9, 1–8 (2018)
11. Kumar, N., et al.: Extremely high magnetoresistance and conductivity in the type-II Weyl
semimetals WP2 and MoP2. Nat. Commun. 8, 1–8 (2017)
12. Shekhar, C., et al.: Extremely large magnetoresistance and ultrahigh mobility in the topolog-
ical Weyl semimetal candidate NbP. Nat. Phys. 11, 645–649 (2015)
13. Bedoya-Pinto, A., Pandeya, A., Liu, D., Deniz, H., Chang, K., Tan, H., Han, H., Jena, J.,
Kostanovskiy, I., Parkin, S.S.P.: ACS Nano. 14, 4405–4413 (2020). https://pubs.acs.org/doi/
10.1021/acsnano.9b09997
14. Cho, A.Y., Arthur, J.R.: Molecular beam epitaxy. Prog. Solid-State Chem. 10, 157–191 (1975)
15. Arthur, J.R.: Molecular beam epitaxy. Surf. Sci. 500, 189–217 (2002)
16. Cho, A.Y.: How molecular beam epitaxy (MBE) began and its projection into the future.
J. Cryst. Growth. 201/202, 1–7 (1999)
17. Klitzing, K.v., Dorda, G., Pepper, M.: New method for high-accuracy determination of the
fine-structure constant based on quantized hall resistance. Phys. Rev. Lett. 45, 494–497
(1980)
18. Tsui, D.C., Stormer, H.L., Gossard, A.C.: Two-dimensional Magnetotransport in the extreme
quantum limit. Phys. Rev. Lett. 48, 1559–1562 (1982)
19. Esaki, L. – Nobel lecture. https://www.nobelprize.org/prizes/physics/1973/esaki/lecture/
20. Alferov. Z.I. – Nobel lecture. https://www.nobelprize.org/prizes/physics/2000/alferov/lecture/
21. Li, Y.-Y., Wang, G., Zhu, X.-G., Liu, M.-H., Ye, C., Chen, X., Wang, Y.-Y., He, K.,
Wang, L.-L., Ma, X.-C., Zhang, H.-J., Dai, X., Fang, Z., Xie, X.-C., Liu, Y., Qi, X.-L., Jia,
J.-F., Zhang, S.-C., Xue, Q.-K.: Intrinsic topological insulator Bi2 Te3 thin films on Si and
their thickness limit. Adv. Mater. 22, 4002–4007 (2010)
22. Zhang, Y., Chang, T.-R., Zhou, B., Cui, Y.-T., Yan, H., Liu, Z., Schmitt, F., Lee, J., Moore,
R., Chen, Y., Lin, H., Jeng, H.-T., Mo, S.-K., Hussain, Z., Bansil, A., Shen, Z.-X.: Direct
observation of the transition from indirect to direct bandgap in atomically thin epitaxial
MoSe2 . Nat. Nanotech. 9, 111–115 (2014)
23. Henini, M. (ed.): Molecular Beam Epitaxy: from Research to Mass Production, p. 13. Elsevier
Science, Waltham (2013)
24. Herman, M.A., Sitter, H.: Molecular Beam Epitaxy: Fundamentals and Current Status.
Springer-Verlag, Berlin/Heidelberg (1989)
25. Herman, M.A., Richter, W., Sitter, H.: Epitaxy: Physical Principles and Technical Implemen-
tation. Springer-Verlag, Berlin/Heidelberg (2004)
26. Cheng, K.-Y.: Molecular beam epitaxy technology of III–V compound semiconductors for
optoelectronic applications. Proc. IEEE. 85, 1694–1714 (1997)
27. Huang, B., et al.: Layer-dependent ferromagnetism in a van der Waals crystal down to the
monolayer limit. Nature. 546, 270–273 (2017)
28. Fei, Z., et al.: Two-dimensional itinerant ferromagnetism in atomically thin Fe3GeTe2. Nat.
Mater. 17, 778–782 (2018)
29. Chen, W., et al.: Direct observation of van der Waals stacking-dependent interlayer mag-
netism. Science. 366, 983–987 (2019)
30. Bedoya-Pinto, A. et al.: Intrinsic 2DXY ferromagnetism in a van der Waals monolayer. https://
arxiv.org/abs/2006.07605
31. Vaz, C.A.F., Bland, J.A.C., Lauhoff, G.: Magnetism in ultrathin film structures. Rep. Prog.
Phys. 71, 056501 (2008)
32. Liu, S., et al.: Wafer-scale two-dimensional ferromagnetic Fe3GeTe2 thin films grown by
molecular beam epitaxy. npj 2D Mater. Appl. 1, 30 (2017)
1198 A. Bedoya-Pinto et al.

33. Hasan, M.Z., Kane, C.L.: Colloquium: topological insulators. Rev. Mod. Phys. 82, 3045–3067
(2010)
34. Kane, C.L., Mele, E.J.: Z2 topological order and the quantum spin Hall effect. Phys. Rev.
Lett. 95, 146802 (2005)
35. Bernevig, B.A., Zhang, S.-C.: Quantum spin Hall effect. Phys. Rev. Lett. 96, 106802 (2006)
36. Bernevig, B.A., Hughes, T.L., Zhang, S.-C.: Quantum spin Hall effect and topological phase
transition in HgTe quantum wells. Science. 314, 1757–1761 (2006)
37. Zhang, H., Liu, C.-X., Qi, X.-L., Dai, X., Fang, Z., Zhang, S.-C.: Topological insulators in
Bi2 Se3 , Bi2 Te3 and Sb2 Te3 with a single Dirac cone on the surface. Nat. Phys. 5, 438–442
(2009)
38. König, M., Wiedmann, S., Brüne, C., Roth, A., Buhmann, H., Molenkamp, L.W., Qi, X.-
L., Zhang, S.-C.: Quantum spin Hall insulator state in HgTe quantum wells. Science. 318,
766–770 (2007)
39. Becker, C.R., Brüne, C., Schäfer, M., Roth, A., Buhmann, H., Molenkamp, L.W.: The
influence of interfaces and the modulation doping technique on the magneto-transport
properties of HgTe based quantum wells. Phys. Status Solidi. 4, 3382–3389 (2007)
40. Cheng, P., Song, C., Zhang, T., Zhang, Y., Wang, Y., Jia, J.-F., Wang, J., Wang, Y., Zhu, B.-F.,
Chen, X., Ma, X., He, K., Wang, L., Dai, X., Fang, Z., Xie, X., Qi, X.-L., Liu, C.-X., Zhang,
S.-C., Xue, Q.-K.: Landau quantization of topological surface states in Bi2 Se3 . Phys. Rev.
Lett. 105, 076801 (2010)
41. Jiang, Y., Wang, Y., Chen, M., Li, Z., Song, C., He, K., Wang, L., Chen, X., Ma, X.,
Xue, Q.-K.: Landau quantization and the thickness limit of topological insulator thin films
of Sb2 Te3 . Phys. Rev. Lett. 108, 016401 (2012)
42. Elliott, S.R., Franz, M.: Colloquium: Majorana fermions in nuclear, particle, and solid-state
physics. Rev. Mod. Phys. 87, 137–163 (2015)
43. Wang, M.-X., Liu, C., Xu, J.-P., Yang, F., Miao, L., Yao, M.-Y., Gao, C.L., Shen, C., Ma, X.,
Chen, X., Xu, Z.-A., Liu, Y., Zhang, S.-C., Qian, D., Jia, J.-F., Xue, Q.-K.: The coexistence of
superconductivity and topological order in the Bi2 Se3 thin films. Science. 336, 52–55 (2012)
44. Xu, J.-P., Liu, C., Wang, M.-X., Ge, J., Liu, Z.-L., Yang, X., Chen, Y., Liu, Y., Xu, Z.-A., Gao,
C.-L., Qian, D., Zhang, F.-C., Jia, J.-F.: Artificial topological superconductor by the proximity
effect. Phys. Rev. Lett. 112, 217001 (2014)
45. Wang, E., Ding, H., Fedorov, A.V., Yao, W., Li, Z., Lv, Y.-F., Xhao, K., Zhang, L.-G.,
Xu, Z., Schneeloch, J., Zhong, R., Ji, S.-H., Wang, L., He, K., Ma, X., Gu, G., Yao, H.,
Xue, Q.-K., Chen, X., Zhou, S.: Fully gapped topological surface states in Bi2 Se3 films
induced by a d-wave high-temperature superconductor. Nat. Phys. 9, 621–625 (2013)
46. Flötotto, D., Ota, Y., Bai, Y., Zhang, C., Okazaki, K., Tsuzuki, A., Hashimoto, T., Eckstein,
J.N., Shin, S., Chiang, T.-C.: Superconducting pairing of topological surface states in bismuth
selenide films on niobium. Sci. Adv. 4, eaar7214 (2018)
47. Chang, C.-Z., Zhang, J., Feng, X., Shen, J., Zhang, Z., Guo, M., Li, K., Ou, Y., Wei, P., Wang,
L.-L., Ji, Z.-Q., Feng, Y., Ji, S., Chen, X., Jia, J., Dai, X., Fang, Z., Zhang, S.-C., He, K., Wang,
Y., Lu, L., Ma, X.-C., Xue, Q.-K.: Experimental observation of the quantum anomalous Hall
effect in a magnetic topological insulator. Science. 340, 167–170 (2013)
48. He, Q.L., Pan, L., Stern, A.L., Burks, E.C., Che, X., Yin, G., Wang, J., Lian, B., Zhou, Q.,
Choi, E.S., Murata, K., Kou, X., Chen, Z., Nie, T., Shao, Q., Fan, Y., Zhang, S.-C., Liu, K.,
Xia, J., Wang, K.L.: Chiral Majorana fermion modes in a quantum anomalous Hall insulator-
superconductor structure. Science. 357, 294–299 (2017)
49. Mellnik, A.R., Lee, J.S., Richardella, A., Grab, J.L., Mintun, P.J., Fischer, M.H., Vaezi, A.,
Manchon, A., Kim, E.-A., Samarth, N., Ralph, D.C.: Spin-transfer torque generated by a
topological insulator. Nature. 511, 449–451 (2014)
50. Khang, N.H.D., Ueda, Y., Hai, P.N.: A conductive topological insulator with large spin Hall
effect for ultralow power spin–orbit torque switching. Nat. Mater.. online published. (2018).
https://doi.org/10.1038/s41563-018-0137-y
51. Fu, L.: Topological crystalline insulators. Phys. Rev. Lett. 106, 106802 (2011)
23 Material Preparation and Thin Film Growth 1199

52. Hsieh, T.H., Lin, H., Liu, J., Duan, W., Bansil, A., Fu, L.: Topological crystalline insulators
in the SnTe material class. Nat. Commun. 3, 982 (2012)
53. Yan, C., Liu, J., Zang, Y., Wang, J., Wang, Z., Wang, P., Zhang, Z.-D., Wang, L., Ma, X., Ji,
S., He, K., Fu, L., Duan, W., Xue, Q.-K., Chen, X.: Experimental observation of Dirac-like
surface states and topological phase transition in Pb1-x Snx Te(111) films. Phys. Rev. Lett. 112,
186801 (2014)
54. Wang, Z., Wang, J., Zang, Y., Zhang, Q., Shi, J.-A., Jiang, T., Gong, Y., Song, C.-L., Ji, S.-H.,
Wang, L.-L., Gu, L., He, K., Duan, W., Ma, X., Chen, X., Xue, Q.-K.: Molecular beam epitaxy
grown SnSe in the rock-salt structure: an artificial topological crystalline insulator material.
Adv. Mater. 27, 4150–4154 (2015)
55. Armitage, N.P., Mele, E.J., Vishwanath, A.: Weyl and Dirac semimetals in three-dimensional
solids. Rev. Mod. Phys. 90, 15001 (2018)
56. Wen, J., Guo, H., Yan, C.-H., Wang, Z.-Y., Chang, K., Deng, P., Zhang, T., Zhang, Z.-D., Ji,
S.-H., Wang, L.-L., He, K., Ma, X.-C., Chen, X., Xue, Q.-K.: Synthesis of semimetal A3 Bi
(A = Na, K) thin films by molecular beam epitaxy. Appl. Surf. Sci. 327, 213–217 (2015)
57. Hellerstedt, J., Edmonds, M., Ramakrishnan, N., Liu, C., Weber, B., Tadich, A.,
O’Donnell, K., Adam, S., Fuhrer, M.: Electronic properties of high-quality epitaxial topo-
logical Dirac semimetal thin films. Nano Lett. 16, 3210–3214 (2016)
58. Schumann, T., Goyal, M., Kim, H., Stemmer, S.: Molecular beam epitaxy of Cd3 As2 on a
III-V substrate. APL Mater. 4, 126110 (2016)
59. Bollinger, A.T., Bozovic, I.: Two-dimensional superconductivity in the cuprates revealed by
atomic-layer-by-layer molecular beam epitaxy. Supercond. Sci. Technol. 29, 103001 (2016)
60. Gozar, A., Logvenov, G., Kourkoutis, L.F., Bollinger, A.T., Giannuzzi, L.A., Muller, D.A.,
Bozovic, I.: High-temperature interface superconductivity between metallic and insulating
copper oxides. Nature. 455, 782–785 (2008)
61. Božović, I., He, X., Wu, J., Bollinger, A.T.: Dependence of the critical temperature in
overdoped copper oxides on superfluid density. Nature. 536, 309–311 (2016)
62. Song, C.-L., Wang, Y.-L., Cheng, P., Jiang, Y.-P., Li, W., Zhang, T., Li, Z., He, K., Wang, L.,
Jia, J.-F., Hung, H.-H., Wu, C., Ma, X., Chen, X., Xue, Q.-K.: Direct observation of nodes
and twofold symmetry in FeSe superconductor. Science. 332, 1410–1413 (2011)
63. Ueda, S., Yamagishi, T., Takeda, S., Agatsuma, S., Takano, S., Mitsuda, A., Naito, M.: MBE
growth of Fe-based superconducting films. Physica C. 471, 1167–1173 (2011)
64. Li, W., Ding, H., Deng, P., Chang, K., Song, C., He, K., Wang, L., Ma, X., Hu, J.-P., Chen, X.,
Xue, Q.-K.: Phase separation and magnetic order in K-doped iron selenide superconductor.
Nat. Phys. 8, 126–130 (2012)
65. Chang, K., Deng, P., Zhang, T., Lin, H.-C., Zhao, K., Ji, S.-H., Wang, L.-L., He, K.,
Ma, X.-C., Chen, X., Xue, Q.-K.: Molecular beam epitaxy growth of superconducting LiFeAs
film on SrTiO3 (001) substrate. Europhys. Lett. 109, 28003 (2015)
66. Wang, Q.-Y., Li, Z., Zhang, W.-H., Zhang, Z.-C., Zhang, J.-S., Li, W., Ding, H., Ou, Y.-
B., Deng, P., Chang, K., Wen, J., Song, C.-L., He, K., Jia, J.-F., Ji, S.-H., Wang, Y.-Y.,
Wang, L.-L., Chen, X., Ma, X.-C., Xue, Q.-K.: Interface-induced high-temperature super-
conductivity in single unit cell FeSe films on SrTiO3 . Chin. Phys. Lett. 29, 037402 (2012)
67. Ge, J.-F., Liu, Z.-L., Liu, C., Gao, C.-L., Qian, D., Xue, Q.-K., Liu, Y., Jia, J.-F.: Supercon-
ductivity above 100 K in single-layer FeSe films on doped SrTiO3 . Nat. Mater. 14, 285–289
(2015)
68. Huang, D., Hoffman, J.E.: Monolayer FeSe on SrTiO3 . Annu. Rev. Condens. Matter Phys. 8,
311–336 (2017)
69. Novoselov, K.S., Geim, A.K., Morozov, S.V., Jiang, D., Zhang, Y., Dubonos, S.V., Grigorieva,
I.V., Firsov, A.A.: Electric field effect in atomically thin carbon films. Science. 306, 666–669
(2004)
70. Xiao, D., Liu, G.-B., Feng, W., Xu, X., Yao, W.: Coupled spin and valley physics in
monolayers of MoS2 and other group-VI Dichalcogenides. Phys. Rev. Lett. 108, 196802
(2012)
1200 A. Bedoya-Pinto et al.

71. Zhang, Y., Ugeda, M.M., Jin, C., Shi, S.-F., Bradley, A.J., Martín-Recio, A., Ryu, H., Kim,
J., Tang, S., Kim, Y., Zhou, B., Hwang, C., Chen, Y., Wang, F., Crommie, M.F., Hussain, Z.,
Shen, Z.-X., Mo, S.-K.: Electronic structure, surface doping, and optical response in epitaxial
WSe2 thin films. Nano Lett. 16, 2485–2491 (2016)
72. Liu, H., Jiao, L., Yang, F., Cai, Y., Wu, X., Ho, W., Gao, C., Jia, J., Wang, N., Fan, H., Yao,
W., Xie, M.: Dense network of one-dimensional midgap metallic modes in monolayer MoSe2
and their spatial undulations. Phys. Rev. Lett. 113, 066105 (2014)
73. Barja, S., Wickenburg, S., Liu, Z.-F., Zhang, Y., Ryu, H., Ugeda, M.M., Hussain, Z., Shen,
Z.-X., Mo, S.-K., Wong, E., Salmeron, M.B., Wang, F., Crommie, M.F., Frank Ogletree,
D., Jeffrey, B., Neaton, A.: Weber-Bargioni. Charge density wave order in 1D mirror twin
boundaries of single-layer MoSe2 . Nat. Phys. 12, 751–756 (2016)
74. Peng, J.-P., Guan, J.-Q., Zhang, H.-M., Song, C.-L., Wang, L., He, K., Xue, Q.-K., Ma, X.-C.:
Molecular beam epitaxy growth and scanning tunneling microscopy study of TiSe2 ultrathin
films. Phys. Rev. B 91, 121113(R) (2015)
75. Ugeda, M.M., Bradley, A.J., Zhang, Y., Onishi, S., Chen, Y., Ruan, W., Ojeda-Aristizabal,
C., Ryu, H., Edmonds, M.T., Tsai, H.-Z., Riss, A., Mo, S.-K., Lee, D., Zettl, A., Hussain,
Z., Shen, Z.-X., Crommie, M.F.: Characterization of collective ground states in single-layer
NbSe2 . Nat. Phys. 12, 92–97 (2016)
76. Ryu, H., Chen, Y., Kim, H., Tsai, H.-Z., Tang, S., Jiang, J., Liou, F., Kahn, S., Jia, C., Omrani,
A.A., Shim, J.H., Hussain, Z., Shen, Z.-X., Kim, K., Min, B.I., Hwang, C., Crommie, M.F.,
Mo, S.-K.: Persistent charge-density-wave order in single-layer TaSe2 . Nano Lett. 18, 689–
694 (2018)
77. Xi, X., Wang, Z., Zhao, W., Park, J.-H., Law, K.T., Berger, H., Forró, L., Shan, J., Mak, K.F.:
Ising pairing in superconducting NbSe2 atomic layers. Nat. Phys. 12, 139–143 (2016)
78. Qian, X., Liu, J., Fu, L., Li, J.: Quantum spin Hall effect in two dimensional transition metal
chalcogenides. Science. 346, 1344–1347 (2014)
79. Tang, S., Zhang, C., Wong, D., Pedramrazi, Z., Tsai, H.-Z., Jia, C., Moritz, B., Claassen, M.,
Ryu, H., Kahn, S., Jiang, J., Yan, H., Hashimoto, M., Lu, D., Moore, R.G., Hwang, C.-C.,
Hwang, C., Hussain, Z., Chen, Y., Ugeda, M.M., Liu, Z., Xie, X., Devereaux, T.P., Crommie,
M.F., Mo, S.-K., Shen, Z.-X.: Quantum spin Hall state in monolayer 1T’-WTe2 . Nat. Phys.
13, 683–687 (2017)
80. Chen, P., Pai, W.W., Chan, Y.-H., Sun, W.-L., Xu, C.-Z., Lin, D.-S., Chou, M.Y., Fedorov,
A.-V., Chiang, T.-C.: Large quantum-spin-Hall gap in single-layer 1T’ WSe2 . Nat. Commun.
9, 2003 (2018)
81. Tang, S., Zhang, C., Jia, C., Ryu, H., Hwang, C., Hashimoto, M., Lu, D., Liu, Z., Devereaux,
T.P., Shen, Z.-X., Mo, S.-K.: Electronic structure of monolayer 1T’-MoTe2 grown by
molecular beam epitaxy. APL Materials. 6, 026601 (2018)
82. Chang, K., Liu, J., Lin, H., Wang, N., Zhao, K., Jin, F., Zhong, Y., Hu, X., Duan, W., Zhang,
Q., Fu, L., Xue, Q.-K., Chen, X., Ji, S.-H.: Discovery of robust in-plane ferroelectricity in
atomic-thick SnTe. Science. 353, 274–278 (2016)
83. Palmstrøm, C.: Epitaxial Heusler alloys: new materials for semiconductor spintronics. MRS
Bull. 28, 725–728 (2003)
84. Hesjedal, T., Ploog, K.H.: Epitaxial Heusler alloys on III-V semiconductors. Handbook of
magnetism and advanced magnetic materials. In: Kronmuller, H., Parkin, S. (eds.) Volume 3:
Novel Techniques for Characterizing and Preparing Samples. Wiley, Hoboken (2007)
85. Li, M.-Y., Shi, Y., Cheng, C.-C., Lu, L.-S., Lin, Y.-C., Tang, H.-L., Tsai, M.-L., Chu, C.-W.,
Wei, K.-H., He, J.-H., Chang, W.-H., Suenaga, K., Li, L.-J.: Epitaxial growth of a monolayer
WSe2 -MoS2 lateral p-n junction with an atomically sharp interface. Science. 349, 524–528
(2015)
86. Bednorz, J.G., Müller, K.A.: Possible highTc superconductivity in the Ba−La−Cu−O
system. Z. Phys. B Condens Matter. 64(2), 189–193 (1986)
87. Dijkkamp, D., Venkatesan, T., Wu, X.D., Shaheen, S.A., Jisrawi, N., Min-Lee, Y.H., McLean,
W.L., Croft, M.: Preparation of Y-Ba-Cu oxide superconductor thin films using pulsed laser
evaporation from high Tc bulk material. Appl. Phys. Lett. 51(8), 619–621 (1987)
23 Material Preparation and Thin Film Growth 1201

88. Singh, R.K., Narayan, J.: Pulsed-laser evaporation technique for deposition of thin films:
physics and theoretical model. Phys. Rev. B. 41(13), 8843–8859 (1990)
89. Amoruso, S., Berardi, V., Bruzzese, R., Spinelli, N., Wang, X.: Kinetic energy distribution of
ions in the laser ablation of copper targets. Appl. Surf. Sci. 127–129, 953–958 (1998)
90. Canulescu, S., Lippert, T., Wokaun, A.: Mass and kinetic energy distribution of the species
generated by laser ablation of La0.6 Ca0.4 MnO3 . Appl. Phys. A. 93(3), 771–778 (2008)
91. Jeong, J., Aetukuri, N., Graf, T., Schladt, T.D., Samant, M.G., Parkin, S.S.P.: Suppression
of metal-insulator transition in VO2 by electric field-induced oxygen vacancy formation.
Science. 339(6126), 1402–1405 (2013)
92. Jeong, J., Aetukuri, N., Passarello, D., Conradson, S.D., Samant, M.G., Parkin, S.S.P.: Giant
reversible, facet-dependent, structural changes in a correlated-electron insulator induced by
ionic liquid gating. PNAS. 112(4), 1013–1018 (2015)
93. Martens, K., Aetukuri, N., Jeong, J., Samant, M.G., Parkin, S.S.P.: Improved metal-insulator-
transition characteristics of ultrathin VO2 epitaxial films by optimized surface preparation of
rutile TiO2 substrates. APL. 104, 081918 (2014)
94. Ohtomo, A., Hwang, H.Y.: A high-mobility electron gas at the LaAlO3/SrTiO3 heterointer-
face. Nature. 427, 423–426 (2004)
95. Chapin, J.S.: Sputtering process and apparatus. Patent 4,166,018, 28 Aug 1979 (Submitted
Jan 31, 1974)
96. Kelly, P.J., Arnell, R.D.: Magnetic sputtering: a review of recent developments and applica-
tion. Vacuum. 56, 159–172 (2000)
97. Brauer, G., et al.: Magnetic sputtering – milestones of 30 years. Vacuum. 84, 1354–1359
(2010)
98. Window, B., Savvides, N.: Unbalanced DC magnetrons as sources of high ion fluxes. J. Vac.
Sci. Techn. A. 4, 453–456 (1986)
99. Musil, J., et al.: Reactive magnetron sputtering of thin films: present status and trends. Thin
Solid Films. 475, 208–218 (2005)
100. Kaufman, H.R., Robinson, R.S.: Operation of Broad-Beam Sources. Commonwealth Scien-
tific Corporation, Alexandria (1987)
101. Child, C.D.: Discharge from hot CaO. Phys. Rev. Series I. 32, 492–511 (1911)
102. Parkin, S.S.P.: Giant magnetoresistance in magnetic nanostructures. Annu. Rev. Mater. Sci.
25, 357–388 (1995)
103. Parkin, S.S.P., et al.: Giant tunneling magnetoresistance at room temperature with MgO (100)
tunnel barriers. Nat. Mater. 3, 862–867 (2004)
104. Yuasa, S., Nagahama, T., Fukushima, A., Suzuki, Y., Ando, K.: Giant room-temperature
magnetoresistance in single-crystal Fe/MgO/Fe magnetic tunnel junctions. Nat. Mater. 3,
868–871 (2004)
105. Grochowski, E.: “The magnetic hard disk drive – today’s technical status and future” at SNIA
Data Storage (DSC) Conference, September 19–22, 2016, Santa Clara Hyatt, Santa Clara, CA

Amilcar Bedoya-Pinto completed his undergraduate studies in


Physics at the Technical University of Munich (TUM) and
received his PhD from the University of Göttingen, where he was
awarded with the Dr. Berliner-Ungewitter Prize (2011). After a
postdoc stay at CiC nanoGUNE (Spain), he moved to Max Planck
Institute of Microstructure Physics, where he currently leads
projects on two-dimensional materials and Weyl semimetals –
grown by MBE – for spintronic applications.
1202 A. Bedoya-Pinto et al.

Kai Chang received his PhD from Tsinghua University in 2015.


He worked at Max Planck Institute of Microstructure Physics
in Germany until 2019 and then joined Beijing Academy of
Quantum Information Sciences in China. He works now on
molecular beam epitaxial growth and in situ scanning tunneling
microscopy studies of low-dimensional quantum materials and
their heterostructures.

Mahesh G. Samant received his PhD from Stanford University in


1986. He has since worked at IBM Research at Almaden Research
Center. His research included structural characterization using
synchrotron radiation-based techniques of materials and inter-
faces, liquid crystal displays, correlated oxides, and spintronics.
He now works on exploratory materials for spin-transfer torque
magnetic random access memory (STT-MRAM).

Stuart Parkin is the Managing Director of the Max Planck


Institute for Microstructure Physics, Halle, Germany, and an
Alexander von Humboldt Professor, Martin Luther University,
Halle-Wittenberg. His research interests include spintronic mate-
rials and devices for advanced sensor, memory, and logic applica-
tions, oxide thin-film heterostructures, topological metals, exotic
superconductors, and cognitive devices. Parkin’s discoveries in
spintronics enabled a more than 10000-fold increase in the storage
capacity of magnetic disk drives. For his work that thereby
enabled the “big data” world of today, Parkin was awarded the
Millennium Technology Award from the Technology Academy
Finland in 2014 (worth 1,000,000 Euro). Parkin is a Fellow/Mem-
ber of: Royal Society (London), Royal Academy of Engineering,
National Academy of Sciences, National Academy of Engineer-
ing, German National Academy of Science – Leopoldina, Royal
Society of Edinburgh, Indian Academy of Sciences, and TWAS –
academy of sciences for the developing world and has received
numerous awards from around the world.
Magnetic Imaging and Microscopy
24
Robert M. Reeve , Hans-Joachim Elmers , Felix Büttner ,
and Mathias Kläui

Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1206
Electron Microscopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1207
Transmission Electron Microscopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1208
Scanning Electron Microscopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1213
Scanning Probe Microscopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1217
Spin-Polarized Scanning Tunneling Microscopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1219
Magnetic Force Microscopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1223
X-Ray Imaging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1227
X-Ray Magnetic Circular Dichroism: A Contrast Mechanism . . . . . . . . . . . . . . . . . . . . . . 1227
TXM: Quick Full-Field Imaging in Transmission Geometry . . . . . . . . . . . . . . . . . . . . . . . 1232
STXM: Optimized for Dynamic Imaging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1233
PEEM: Imaging Surfaces of Bulk Samples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1233
CXI: Zero Drift and Femtosecond Temporal Resolution . . . . . . . . . . . . . . . . . . . . . . . . . . 1235
SP-ARPES: Microscopy in Momentum Space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1237
Medical Magnetic Imaging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1238
Magnetic Resonance Imaging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1238
Nuclear Magnetic Resonance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1239
Nuclear Quadrupole Resonance Imaging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1243
Magnetoencephalography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1244
Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1246
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1246

R. M. Reeve · H.-J. Elmers · M. Kläui ()


Institute of Physics, Johannes Gutenberg University Mainz, Mainz, Germany
e-mail: reeve@uni-mainz.de; elmers@uni-mainz.de; klaeui@uni-mainz.de; mathias@klaeui.de
F. Büttner
Helmholtz-Zentrum Berlin für Materialien und Energie, Berlin, Germany
e-mail: felix.buettner@helmholtz-berlin.de

© Springer Nature Switzerland AG 2021 1203


J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_28
1204 R. M. Reeve et al.

Abstract

The magnetic domain configuration of a system reveals a wealth of information


about the fundamental magnetic properties of that system and can be a critical
factor in the operation of magnetic devices. Not only are the details of the
domain structure strongly governed by materials parameters, but in thin-film
and mesoscopic elements, the geometry has an often pivotal effect, providing
a convenient handle to tailor desired domain states. Furthermore, a full under-
standing of a system requires, in addition, investigation of the dynamic evolution
of the spin state, which is of particular importance for applications relying on,
e.g., the switching of magnetic elements. In this chapter, we review some of
the main modern techniques for magnetic imaging, highlighting their respective
advantages and limitations. The methods for imaging domain configurations
and spin structures cover various spatial and temporal resolution scales and
encompass those based on electron and x-ray microscopy as well as scanning
probe techniques. Furthermore, away from the discipline of condensed-matter
physics, magnetic effects are instrumental in a number of techniques for medical
imaging, some key examples of which we also present.

List of Abbreviations

APD Avalanche photodiode


ARPES Angle-resolved photoemission spectroscopy
BOLD Blood-oxygen-level dependent
CCD Charge-coupled device
CXI Coherent x-ray imaging
EEG Electroencephalography
FEL Free electron laser
fMRI Functional magnetic resonance imaging
FTH Fourier transform holography
KZP Condenser zone plate
LUMO Lowest unoccupied molecular orbital
MFM Magnetic force microscopy
MEG Magnetoencephalography
MRFM Magnetic resonance force microscopy
MRI Magnetic resonance imaging
MZP Objective zone plate lens
NMR Nuclear magnetic resonance
NQR Nuclear quadrupole resonance
OSA Order selecting aperture
PEEM Photoemission electron microscopy
RF Radio frequency
SP-ARPES Spin-polarized angle-resolved photoemission spectroscopy
SP-STM Spin-polarized scanning tunneling microscopy
24 Magnetic Imaging and Microscopy 1205

SEM Scanning electron microscope


SEMPA Scanning electron microscopy with polarization analysis
SPLEED Spin-polarized low-energy electron diffraction
SQUID Superconducting quantum interference device
STM Scanning tunneling microscopy
STXM Scanning transmission x-ray microscopy
TEM Transmission electron microscopy
TXM Transmission x-ray microscopy
UHV Ultrahigh vacuum
XMCD X-ray magnetic circular dichroism

List of Symbols

V Bias voltage
N Counts
τ Delay time
D() Density of states
δ Dirac delta function
Seff Effective Sherman/sensitivity function
e Electron charge
ε Energy
M0 Equilibrium magnetization
c∗ Experimental constant
εF Fermi energy
f (ε) Fermi function
f Frequency
dFm
dz Gradient of the magnetic force
γ Gyromagnetic ratio
m, n Integers
wL Larmor angular frequency
flor Lorentz force
B Magnetic flux density
M Magnetization
 Orbital quantum number
ml Orbital angular momentum magnetic quantum number
q Orbital angular momentum of photon
r0 Origin position vector
r Position vector
Q Quality factor
T1,2 Relaxation times
f0 Resonance frequency
PS Sample spin polarization
s Spin quantum number
A Spin asymmetry
1206 R. M. Reeve et al.

ρS Spin-dependent sample density of states


ρT Spin-dependent tip density of states
ms Spin magnetic quantum number
P Spin polarization
PT Tip polarization
φ Tip-sample magnetization angle
j Total angular momentum
mj Total angular momentum magnetic quantum number
I Tunneling current
M Tunneling matrix element
g0 Unpolarized conductivity
υ Velocity

Introduction

The magnetic domain configuration of a system reveals a wealth of information


about the fundamental magnetic properties of that system and can be a critical
factor in the operation of magnetic devices [1]. Not only are the details of the
domain structure strongly governed by materials parameters, but in thin film and
mesoscopic elements, the additional contributions of shape and configurational
anisotropy mean that the spin configurations are qualitatively different from bulk
systems and hence geometrical control provides a convenient handle to tailor desired
domain states [2]. Furthermore, more recently, it is not only the magnetic domain
patterns which are of interest but also the spin configurations of the magnetic domain
walls themselves since these have been proposed as functional elements in next-
generation memory, logic, and sensing devices [3, 4, 5] and such quasiparticle-like
spin textures interact differently with magnetic fields and currents depending on the
detailed spin structure [6, 7]. Hence, high spatial resolution imaging techniques
are becoming increasingly important. A full understanding of a system requires
in addition investigation of the dynamic evolution of the spin state, which is of
particular importance for applications relying on, for example, the switching of
magnetic elements. Since the dynamics of such systems are governed by precession
frequencies and are then typically in the GHz regime, high temporal resolution time
domain imaging is also highly desirable. Away from the discipline of condensed-
matter physics, magnetic effects are instrumental in a number of techniques for
medical imaging where the requirements and desired attributes of the methods are
very different.
One of the first direct observations of magnetic domain structure was achieved in
1932 via the Bitter technique [8]. In this method, the sample under investigation is
covered with a fluid containing a suspension of ferromagnetic particles. Depending
on the size and properties of the suspension and the sample, the particles are found
to align in the magnetic stray field from the sample, and the resulting pattern can
be imaged with conventional optical or electron microscopy. While this technique
remains in use, in the intervening years, a whole range of other imaging techniques
have been developed. Given the wide range of length and timescales that can be
24 Magnetic Imaging and Microscopy 1207

relevant and due to the large range of systems that it is possible to image, there is no
universal best technique, and it is necessary to carefully select the most appropriate
approach based on the particular requirements of a given application or experiment.
Furthermore, since different techniques are sensitive to different magnetic properties
such as the stray field or magnetization, it can be best to combine multiple options
for a more comprehensive understanding.
In this chapter, we review some of the main techniques currently employed for
magnetic imaging. The aim is not to provide an exhaustive list of techniques but
rather to give an overview of some of the most widely employed options, high-
lighting some of the particular considerations that must be taken into account when
selecting an appropriate method, with the particular advantages and limitations of
the techniques highlighted. References are provided to more in-depth discussions
of particular techniques. Reviews and comparisons of multiple techniques are
provided in [1, 9, 10, 11]. Furthermore, we primarily focus on the application of
techniques to ferromagnetic systems. However, with the recent intense interest in
antiferromagnetic spintronics, there have been a number of advances in the field
of imaging of antiferromagnetic domains, as reviewed in Ref. [12]. We divide the
methods into different categories based on either the nature of the probing radiation,
such as electron beam or x-ray illumination, or in the case of the scanning probe
methods based on the principle of operation. While optical techniques such as Kerr
microscopy are also very widely applied to image domain structures and other spin
states such as magnetic skyrmion bubbles, we do not cover these in detail here
but rather refer the reader to the dedicated chapters on magneto-optic effects and
literature reviews of recent advances in the field [13, 14, 15]. Such optical methods
are generally more limited in their spatial resolution than many of techniques
presented in this chapter, yet they offer a number of advantages including being
relatively quick, inexpensive, and easy to apply. Finally, at the end of the chapter, we
briefly introduce some of the key modern magnetic imaging techniques for medical
applications.

Electron Microscopy

The first class of microscopy techniques that will be discussed uses electron beams
in order to probe the sample. The incident electron beam can interact with the
sample based on different mechanisms. In the first instance, the beam may be
deflected depending on the magnetic configuration, yet additionally, the excitation
of the sample generates secondary electrons which also carry information about
the magnetic state. Due to the mature technology in generating highly focused
electron beams, scanning approaches can offer very good spatial resolution imaging.
Electron microscopy techniques tend to be limited to conducting specimens, since
otherwise the sample becomes charged by the electron beam, leading to unwanted
deflections and distortions. This can sometimes be overcome if insulating systems
want to be investigated by coating the surface with a thin conductive layer to help
mitigate charge buildup. Furthermore, the application of magnetic fields during
imaging is often severely limited due to their deflecting and depolarizing effects
1208 R. M. Reeve et al.

on the electrons, although there are sometimes strategies to partially overcome


this [16]. Here, we focus on methods which use an unpolarized electron beam as the
probe; however, we note that certain specialized approaches such as spin-polarized
low-energy electron microscopy employ polarized electron beams to excite the
sample, details of which can be found in [9, 17, 18].

Transmission Electron Microscopy

Following the Bitter technique, transmission electron microscopy (TEM) was one
of the earliest techniques for revealing magnetic domain structure [19]. In TEM
imaging, the electron beam incident on the specimen is accelerated by high voltages,
resulting in highly energetic electrons with typical energies of 100–300 keV and in
some cases up to 1000 keV. The electron intensity is then detected in transmission.
One key advantage of TEM is the ease of carrying out complementary non-
magnetic characterization of samples in order to correlate the observed magnetic
configurations with the local electronic and structural properties. However, since
the signal is measured in transmission, this places considerable constraints on the
sample thickness. For scanning TEM, sample thicknesses are typically limited to
about 100 nm or less, whereas for Lorentz TEM, thicknesses up to around ∼300 nm
are possible. For bulk samples, it is necessary to apply thinning processes before
imaging, which potentially modifies the domain structure of interest. Thin-film
samples and associated lithographically defined nanostructures are more readily
imaged via TEM, but need to be deposited on suitable substrates which are
transmissive for the electrons, e.g., silicon nitride (Si3 N4 ) membranes, which take
some care in handling. Another important consideration when applying TEM
imaging to magnetic structures is that usually the sample would be subject to a
very strong magnetic field from the objective lens of the microscope. To avoid
perturbing the domain structure, strategies have to be employed to reduce this
field which may require dedicated equipment and in all cases tend to limit the
resolution of magnetic modes of TEM microscopy as compared to other forms
of TEM characterization. The existence of structural contrast even in magnetic
imaging modes can also limit the practical resolution. A review of the application
of TEM to imaging magnetic microstructure is provided in [20]. In the following,
we describe some of the different operational modes of the technique. These modes
have different advantages and drawbacks and are sensitive to varying features of
the magnetic state of the system. However, in general, a number of modes can be
combined to provide more comprehensive information and overcome the limitations
of a particular mode [21].

Lorentz Microscopy
Lorentz microscopy relies on the perturbation of an electron beam due to magnetic
fields. The resulting small angular deflections of the beam of around 10−5 −
−10−4 rad, can be classically attributed to the so-called Lorentz force:
24 Magnetic Imaging and Microscopy 1209

flor = |e| (υ × B) , (1)

where e is the electron charge; υ is the electron velocity, which depends on the
acceleration energy; and B is the magnetic flux density. In quantum mechanical
terms, the sample can be considered to modulate the phase of the incident electron
wave depending on the magnetic state, leading to bright-dark contrast due to
interference. From (1), it can be seen that components of B aligned with the beam
do not contribute to the deflection and therefore samples may need to be tilted in
the case of perpendicular magnetic anisotropy systems. The technique is sensitive
to magnetic flux along the whole path of the electron beam, and hence, it is not only
the magnetization within the sample which contributes to the deflection but also
stray magnetic fields. In some cases, these two contributions can act against each
other, diminishing or even cancelling out contrast. The deflection direction depends
on the magnetization orientation and is perpendicular to it. The magnitude of the
deflection is directly proportional to the thickness and averaged magnetic induction
of the specimen.

Fresnel Imaging
The first mode of Lorentz imaging is the Fresnel imaging, or defocus mode. Since
the influence of the magnetic structure of the specimen only causes deflections of
the beam, an in-focus image of the sample normally does not contain any magnetic
contrast. In Fresnel imaging, this is overcome by defocusing the imaging lens. This
reveals magnetic features of the sample, however, at the expense of reducing the
achievable spatial resolution. For this imaging mode, the objective lens is kept
switched off to avoid the exposure of the specimen to high magnetic fields from
the lens which would result in a destruction of domain structures of interest. For
this reason, a Lorentz mini-lens is used to image the magnetic structure, and the
objective lens can be used to produce a vertical magnetic field.
The mechanism is illustrated in Fig. 1. As can be seen in the figure, the opposite
direction of deflection for neighboring beams on either side of a 180◦ domain wall
leads to either converging or diverging beams at the wall positions. Consequently,
such walls are revealed by corresponding bright or dark contrast, depending on
whether overfocused or underfocused imaging is employed, while the domains
themselves usually have uniform contrast. An exception is the case of polycrys-
talline films where due to small fluctuations in the directions of magnetocrystalline
anisotropy characteristic ripple contrast can occur which is oriented perpendicularly
to the magnetization direction of the given domain. More in-depth reconstructions
of the domain states of a sample are, however, extremely challenging in this imaging
mode. Due to the required high defocusing, there is a very strong nonlinearity
between the contrast and the magnetic state of the film. More recently, the possibility
to reconstruct the phase of the emerging electron wave has been demonstrated by
acquiring multiple Fresnel images for different values of defocus and then applying
the transport of intensity equations [22].
1210 R. M. Reeve et al.

Fig. 1 Fresnel mode of Lorentz microscopy: Schematic representation of the contrast formation in
the Fresnel imaging mode for a sample consisting of two opposing in-plane domains separated by
a 180◦ domain wall. The Lorentz force deflects the electron beams, leading to dark/bright contrast
for over- and underfocus conditions of the microscope, respectively

Foucault Imaging
In Foucault imaging, a different strategy is employed to reveal magnetic contrast in
TEM images [23]. Since the variously magnetized domains lead to different Lorentz
deflection angles of the beams, the reciprocal space image of the sample is split into
different components for these different Lorentz angles. Particular components can
therefore be selected by using an aperture to block part of this reciprocal space
pattern which is formed in the back focal plane of the imaging lens and blocking
aperture. This is illustrated schematically in Fig. 2 where two spatially separated
diffraction spots are evident due to the two magnetization directions present in the
specimen. By blocking one or other of these beams, contrast is generated in the
image. Unlike in Fresnel imaging, in the Foucault mode, the contrast is correlated
to the magnetic induction from the domains themselves and not the change in
magnetic induction between domains, resulting in an image that is easy to interpret.
Furthermore, the magnetic imaging is performed by keeping the imaging lens in
focus and thereby generating better spatial resolution of the magnetic structure as
compared to the Fresnel mode, which is usually performed with a high defocus of
the imaging lens. However, the stringent requirements on the quality and positioning
of the aperture mean that this mode is more difficult to implement.

Differential Phase Contrast Microscopy


The modes discussed so far are attractive since they are relatively easy to implement
in an existing microscope and it is relatively easy to perform the measurements
and interpret the data. However, the requirement for defocused images limits
24 Magnetic Imaging and Microscopy 1211

Fig. 2 Foucault mode of


Lorentz microscopy:
Schematic representation of
the contrast formation in the
Foucault imaging mode for a
sample consisting of two
opposing in-plane domains
separated by a 180◦ domain
wall. Part of the diffraction
plane is blocked, leading to
bright-dark contrast
corresponding to the domain
regions

the achievable resolution for Fresnel imaging, and both techniques offer limited
quantitative information. For more quantitative imaging, the differential phase
contrast technique is an attractive option based on scanning TEM [24, 25]. The
incident focused beam is rastered across the sample and the transmitted beam
detected by a special four-quadrant circular detector. For magnetic samples, the
Lorentz deflection leads to opposite quadrants being illuminated to a greater or
lesser extent, and hence, difference signals for the two opposing quadrant pairs
provide quantitative information concerning the two orthogonal in-plane magnetic
induction components. This is illustrated schematically in Fig. 3.
This method suffers from longer recording times than the previous modes due
to the necessity to scan the sample and furthermore has increased instrumental
and experimental complexity. Nevertheless, good spatial resolution can be achieved
from the focused beam, down to around 5 nm. One drawback is that parasitic
differential contrast of non-magnetic structural origin can also arise, in particular
for thinner samples and polycrystalline systems. However, the use of fully pixelated
detectors is a recent advance which provides much improved efficiency for enhanced
magnetic contrast [26].

Electron Holography
Whereas the modes discussed so far can conceptually be understood from a classical
picture of the electron beam, electron holography explicitly relies on the quantum
mechanical wave nature of the electrons. Holographic techniques can provide high
spatial resolution imaging, typically down to about 5 nm. Furthermore, they offer the
possibility to image stray fields outside a given sample. A wide variety of schemes
exist, including even tomography, as reviewed in [27, 28, 29]; however, the most
common mode is off-axis holography as outlined below.
1212 R. M. Reeve et al.

Fig. 3 Differential phase contrast microscopy with a four-quadrant detector: Depending on the
magnetic induction in the sample, the beam is deflected in the perpendicular direction. The
difference signal between opposite quadrants of the detector provides a differential phase contrast.
For the first case, the beam is aligned with the magnetic induction, and there is no deflection,
yielding an equal signal for the left (L) and right (R) quadrants. In the second case, the beam
passes through a domain oriented out of the plane of the page, leading to a right deflection of the
beam and L<R. In the final case, the beam passes through a domain oriented into the plane, leading
to the opposite deflection of the beam and correspondingly L>R

Off-Axis Holography
In off-axis holography, a highly coherent incident beam is split into a probe beam
and a reference beam, the first of which passes through the sample, while the latter
remains unperturbed. Due to the interaction of the probe beam with the magnetic
state, the electron wave acquires a phase shift depending on its path. When the
probe and reference beams are then recombined, they interfere to form a holographic
interference image, encoding information on both the phase and amplitude of
the transmitted wave. In the ideal case, this is directly related to the magnetic
state of the sample; however, complications arise for samples with nonuniform
composition or thickness since these can introduce other sources of phase shifts, for
instance electrical. Quantitative information can then be extracted by processing the
interference pattern to mathematically reconstruct the amplitude and phase. There
are two main imaging modes. The setup for the absolute mode is depicted in Fig. 4.
The sample is chosen such that it only partly fills the image plane, for example by
imaging the edge of a structure or a small element. Part of the beam then passes
through the specimen and part is unperturbed. In order to recombine the two beams
and form the hologram, an element called a biprism is used which consists of a thin
metallic wire or quartz fiber coated with Au or Pt and which is biased at a voltage
of typically 50–200 V. In the differential mode, two beams are created which are
both directed toward the sample, separated by a small distance, and the phase shift
between these two beams is then recorded. This imaging technique is advantageous
24 Magnetic Imaging and Microscopy 1213

Fig. 4 Electron holography:


Schematic representation of
the off-axis electron
holography technique

for the investigation of fine structure such as the profiles of magnetic domain walls
with the resolution set by the beam diameter. It does, however, require a specially
adapted microscope and image reconstruction software [21].

Aberration Correction
In conventional TEM, there has been significant recent improvement in achievable
resolution by implementing aberration correction [30]. Electron lenses are inher-
ently much poorer than optical lenses, and their associated spherical aberration is
often a key limiting factor in determining the resolution of electron microscopes.
To counteract this, schemes have been developed to compensate for the aberration
of the TEM objective lens by incorporating a correcting element with negative
spherical aberration into the microscope. Two approaches are based on multipole
lenses called the quadrupole-octupole corrector and sextupole corrector. In the case
of magnetic imaging, the objective lens is often not used due to the impact of the
associated magnetic field on the sample; however, aberration correction schemes can
still be employed to compensate for the relevant lens in the instrument and improve
the attainable resolution of these imaging modes [31].

Scanning Electron Microscopy

In a scanning electron microscope (SEM), the electron beam is scanned across the
sample, and the generated electrons are detected in reflection. When an energetic
primary electron beam interacts with a sample, a spectrum of energies for electrons
leaving the sample results. At high energies around the primary beam energy, there
is a peak corresponding to elastically scattered electrons. In the middle of the spec-
trum, small elemental specific peaks are found corresponding to Auger electrons
1214 R. M. Reeve et al.

which are typically in the 100–2000 eV range. Finally, at very low energies, below
around 50 eV, one finds the so-called true secondary electrons. These correspond to
electrons having undergone many inelastic scattering processes which are emitted
in a cascade process. Since this scattering involves states in the vicinity of the
Fermi level, the emitted secondaries are found to be spin-polarized in itinerant
ferromagnets due to the imbalance between spin-up and spin-down states [32]. For
energies above ∼10 eV, this spin polarization directly reflects that at the Fermi level,
whereas at lowest energies, an additional enhancement in polarization is observed
due to a spin-dependent scattering-induced spin-filtering effect. The interaction of
the electrons with the sample leads to different effects depending on the energy,
and these can variously be employed for magnetic imaging [33]. In the first case,
as with many of the TEM techniques, the trajectories of the detected electrons
are modified by the magnetic configuration of the sample. In the second case,
the emitted electrons may be spin-polarized, with the polarization representing
features of the spin-split band structure of the ferromagnet. If the low-energy
secondary electrons are detected, the contrast is sometimes termed “type I.” In
this case, deflection is largely due to the magnetic stray fields from the sample.
The elastically backscattered electrons are primarily affected by the magnetization
within the sample, leading to so-called “type II” contrast. In this case, the sample is
tilted with respect to the beam, and the resulting deflection of the electrons within
the sample leads to an enhancement or reduction of the backscattered electron
yield depending on whether the deflection is directed toward or away from the
surface. While the lateral resolution, at a few μm, is quite limited compared to other
magnetic imaging techniques, the deep penetration of highly energetic electrons
can be taken advantage of to image through surface layers and can probe domain
structures to depths of ∼1–20 μm depending on the incident electron beam energy.

SEMPA
Scanning electron microscopy with polarization analysis (SEMPA) or spin-SEM
takes advantage of the fact that the spin polarization of the emitted low-energy
electrons is oriented antiparallel to the magnetization in the sample [34, 35, 36].
Hence, by exciting secondary electrons point-by-point with an unpolarized scanning
electron beam and measuring the polarization of the emitted electrons, a direct
quantitative representation of the domain state can be obtained as illustrated in
Fig. 5. The polarization of the emitted electrons is measured via a spin detector.
Here, the spin-polarized electrons are focused onto a target where, due to the
spin-orbit interaction, asymmetries in scattering exist for spin-up and spin-down
electrons. By counting the number of electrons scattered in opposing directions
using electron multipliers, the beam polarization can be calculated as follows:

1 NA − NB
P = , (2)
Seff NA + NB

where Seff is the effective Sherman or sensitivity factor of the detector which
quantifies the scattering asymmetry that is obtained for a 100% polarized beam
24 Magnetic Imaging and Microscopy 1215

Fig. 5 SEMPA: Schematic representation of the SEMPA technique. An unpolarized SEM beam
excites secondary electrons from the magnetic sample, and their spin polarization is analyzed with
a spin detector. On the right, typical images are presented for an Fe whisker. Here, the technique
simultaneously acquires the topography and the two in-plane magnetization components which
can be processed to provide the full in-plane color map of the magnetization

and NA/B are the electron counts for scattering in opposite directions. Instruments
are usually equipped with two pairs of electron counters which simultaneously
provide the two orthogonal in-plane components of the magnetization which can
be combined into a single map of the 2D domain structure, as shown in Fig. 5.
Some instruments also employ a spin-rotator, in which case the out-of-plane
component can also be accessed. The sum of the signals from all four detectors
provides a secondary electron topographic image of the sample which is helpful
in distinguishing between features of magnetic and non-magnetic origin. One
advantage of SEMPA is that in general morphological details are suppressed in the
magnetic asymmetry images.
A number of designs of spin detector exist [37]; however, the two most
commonly employed in SEMPA are the Mott polarimeter and spin-polarized
low-energy electron diffraction (SPLEED) detector. The SPLEED detector takes
advantage of the spin-dependent low-energy electron diffraction from a W(100)
crystal with the asymmetries in the intensities of the (2,0) diffraction beams at
104.5 eV scattering energy employed [38]. The Mott detector is based on the spin-
dependent Mott scattering of highly energetic electrons from films of high atomic
number elements. Originally, Mott polarimeters worked at particularly high voltages
and were very bulky. Nowadays, however, much more compact instruments are
available operating around 25 kV [39], facilitating their employment in a small lab
setting [40]. Unfortunately, the inherent low efficiency of spin detectors of around
10−4 means that long acquisition times are required to obtain sufficient signal-
to-noise ratio per pixel, with typical images requiring several minutes or longer,
depending on the desired resolution, the imaging area, and the particular material.
The entire system therefore needs to be stable on these timescales including the
1216 R. M. Reeve et al.

incident beam, sample state, and mechanical vibrations. Furthermore, thermal drifts
are often problematic and can limit practical resolutions. In the case of the SPLEED
detector, the integrity of the W surface also needs to be maintained, since even
residual gas adsorption will degrade the performance [41], requiring that the surface
is periodically flash-heated to regenerate good scattering conditions.
Since the magnetic probing depth of spin-polarized electron spectroscopies is
very small [42], SEMPA has an extreme surface sensitivity of around 1 nm. On
the one hand, this places particularly stringent requirements on the cleanliness of
surfaces being measured, requiring measurement in ultrahigh vacuum (UHV). For
thin films, capping layers or non-magnetic oxide surface regions often need to be
removed in situ by, e.g., Ar+ ion sputtering, while bulk crystals are often cleaved in
situ to reveal a pristine surface. For simple 3d metals, such in situ sputtering often
yields good results, yet for more complex compounds, care needs to be taken to
ensure that different sputter rates for the different elemental components do not lead
to changes in the stoichiometry. Recently, however, it has been demonstrated that
it is possible to image through heavy metal capping layers provided they remain
thin enough [43]. An often employed strategy to mitigate the surface requirements
involves the in situ deposition of a thin dusting layer of Fe [44]. The expectation
is that this will couple to the magnetic structure of the underlying film, imprinting
the domain structure in the pristine Fe layer which can itself then be imaged. This
method can also be employed to improve the contrast for materials with low signals,
such as non-itinerant ferromagnets, and can also facilitate the imaging of insulating
systems where charging effects usually prevent investigation with electron beams.
However, it is necessary to confirm that the thickness of the deposited Fe film is thin
enough that it does not change the domain state in the material under investigation.
At the same time, the surface sensitivity confers the ability to selectively probe
the properties of the surface [45], which are often instrumental in determining
device operation and which have furthermore been taken advantage of to image
particularly thin ferromagnetic films down to just a few monolayers [46] and even
layered antiferromagnets due to the strong contribution of the uppermost atomic
layer [47].
Virtually, all SEMPA imaging to date has been static or quasi-static in nature.
The long acquisition times are a barrier to investigation of magnetization dynamics;
however, very recently, the feasibility of imaging on nanosecond timescales with
advanced signal processing based on the time of detection of the individual electron
counts has been demonstrated [48, 49]. Via careful data analysis, this also opens up
opportunities for improvements in the signal-to-noise ratio, as well as the detection
of competing magnetic switching pathways that are not accessible in conventional
pump-probe imaging [49, 50]. The high spatial resolution is a key advantage of the
technique, enabling imaging not only of domain configurations but also of domain
wall spin structures [51]. The impetus to increase the spin signal means that it
is usual to operate the SEM at large beam currents and low voltages of typically
1–3 kV, for which the emission of the spin-polarized low-energy secondaries is
increased [52]. Under these conditions, typical resolutions are around 20 nm;
however, resolution of better than 5 nm has been demonstrated [34].
24 Magnetic Imaging and Microscopy 1217

Scanning Probe Microscopy

The increasingly low dimensions of magnetic nanostructures and accordingly the


size of the magnetic structures, e.g., skyrmions, vortices, and domain configura-
tions in general, currently reaching the range of 10 nm [9], need advanced high
spatial resolution microscopy techniques. The scanning probe methods, i.e., spin-
polarized scanning tunneling microscopy (SP-STM) and magnetic force microscopy
(MFM), may have lateral resolutions down to the atomic dimensions, providing
a considerable advantage over many imaging methods. However, the very high
lateral resolution comes at a cost because control of the sample environment and
tip preparation causes considerable additional experimental effort. We begin with
an overview of different scanning probe techniques and then discuss spin-polarized
scanning tunneling microscopy and magnetic force microscopy as the two most
common methods in more detail in the separate sections.
Spin-polarized scanning tunneling microscopy goes down to the utmost lateral
resolution of scanning tunneling microscopy, being able to even resolve electronic
orbitals smaller than atomic distances [53, 54]. The magnetic contrast is introduced
by using a ferromagnetic or antiferromagnetic tip, exploiting the spin-dependent
differences of the density of states in tip and sample. First, results of spin-polarized
scanning tunneling microscopy were already shown in the early 1990s [55], but it
took some years until it became an established method [56, 57]. The difference in
tunneling conductivity is in principle similar to the effect exploited in a magnetic
tunneling magnetoresistance device. The more states that are available to tunnel,
the higher the resulting tunneling current. The tunneling current from the tip is
spin-polarized because of the imbalance of electrons with spin-up and spin-down.
The current is also proportional to the number of free states that are available
for the electrons to tunnel into. Consequently, the current is different for parallel
and antiparallel orientations of tip and sample magnetization. Considering atomic
resolution, the method can also be applied if one or both surfaces of tip and
sample are antiferromagnets avoiding the problem of tip-sample interaction. Spin-
polarized scanning tunneling microscopy definitely requires UHV conditions and
ultraclean surfaces. Although results obtained at room temperature have been
reported, low-temperature experiments considerably increase the mechanical and
electronic stability of the measurement.
Magnetic force microscopy is in principle very well suited to image magnetic
domains with high resolution in an ambient environment. However, it is intrinsically
limited in lateral resolution by the physical effect used to obtain magnetic contrast.
Magnetic force microscopy relies on the long-range magnetic dipolar interaction
of a magnetic tip and the stray field of the sample. The magnetic tip has to
be lifted a few nm above the surface to avoid van der Waals interactions, thus
decreasing the attainable resolution, while the measurable magnetic force rapidly
decreases with distance. This trade-off between resolution and signal limits the
obtainable resolution, and atomic resolution cannot be achieved [57]. Another
drawback is the fact that the tip-sample interaction may easily change the magnetic
structure during scanning. A considerable improvement has been achieved by
1218 R. M. Reeve et al.

exploiting the extremely short-ranged magnetic exchange interaction, as proposed


and experimentally shown for a prototypical antiferromagnetic material NiO [58].
By using atomic force microscopy with a magnetic tip, one detects the short-range
magnetic exchange force between tip and sample spins, revealing the arrangement of
both surface atoms and their spins simultaneously. With this technique, the inter-spin
interactions can be investigated at the atomic level. Since the exchange interaction
is strongly modulated by any material between tip and sample, this method only
works for very clean surfaces in UHV.
Instead of measuring the force due to the dipolar interaction, one can alternatively
measure the stray field directly using a Hall probe [59, 60, 61] or a superconducting
quantum interference device (SQUID) [62]. Both methods are passive measure-
ments avoiding magnetic perturbation of the specimen. This advantage comes at
the cost of less spatial resolution (≈0.35 μm) as determined by the dimensions
of the lithographically fabricated probe. Scanning Hall probe microscopy has the
advantage of a wider operating temperature range and a decent field sensitivity
of 0.1 G. The SQUID probe operated at low temperatures on the other hand is
considerably more sensitive to small fields (10−6 G). A spatial resolution of 10 μm
has been demonstrated [63].
A very special type of magnetic force microscopy is given by magnetic resonance
force microscopy (MRFM). In this experiment, a force signal is generated by modu-
lating the sample magnetization with standard magnetic resonance techniques [64].
The magnetic tip at the end of a cantilever is positioned roughly 100 nm above the
sample surface. The isosurface of constant stray field of the tip defines a resonant
slice representing those points in the sample where the field matches the condition
for magnetic resonance. As the cantilever vibrates, the resonant slice swings back
and forth through the sample causing cyclic adiabatic inversion of the spin. The
cyclic spin inversion causes a slight shift of the cantilever frequency owing to the
magnetic force exerted by the spin on the tip. Spins as deep as 100 nm below
the sample surface can be probed. By moving the tip in all three dimensions, a
tomographic image of the spin distribution can be mapped. The main advantage is
the outstanding sensitivity of this method, providing single electron spin detection
in combination with high spatial resolution of 10 nm [65]. The sensitivity is large
enough to sense nuclear spins, too. Measuring the nuclear spin-lattice relaxation
times locally in the mK temperature range allows the characterization of magnetic
properties of inhomogeneous electron systems realized in oxide interfaces, topo-
logical insulators, and other strongly correlated electron systems such as high-Tc
superconductors [66].
Finally, nitrogen-vacancy (NV) center magnetometry is a very promising emerg-
ing technique which has also been incorporated into atomic force microscopes to
provide particularly sensitive, high spatial resolution magnetic imaging [67, 68].
The approach is based on the proposal by Chernobrod and Berman to use single
electronic spins as local magnetic field sensors [69]. The working principle is
shown in Fig. 8. While the spin is scanned over the surface, the local magnetic field
causes a Zeeman splitting of the spin sublevels, which can be detected optically
24 Magnetic Imaging and Microscopy 1219

by measuring the photoluminescence of the probe in an electron spin resonance


measurement. The probe spin system of choice is a single nitrogen-vacancy defect
in diamond which exhibits the required properties for the measurement including
favorably long coherence times. The nitrogen-vacancy defects are created near the
surface of diamond nanocrystals or nanopillars via high-energy electron/proton
irradiation, followed by annealing. The diamond is then mounted into an AFM
to act as the probe-tip [70]. As the tip is scanned across the surface, microwave
fields are applied to stimulate electronic transitions between the spin triplet sublevels
of the system. On resonance, the photoluminescence spectra show a characteristic
drop in intensity. Due to the Zeeman effect, this feature is split and shifted in
an applied magnetic field, providing a measurement of the projection of the field
along the nitrogen-defect quantization axis that is localized at the defect site.
In this manner, the stray magnetic field from a magnetic vortex core [71] and
even a single electron spin have been imaged [72]. In addition to this high
sensitivity, for the imaging of spin textures, it can provide an excellent spatial
resolution of typically a few 10 s of nm, depending on the sample surface to
nitrogen-vacancy defect separation. It is also non-invasive and has successfully
been employed to image the pinning and propagation of magnetic domain walls in
nanowires [73] without undesired perturbations of the magnetic state from the tip.
Even the reconstruction of the full 3D spin texture is possible under the minimum
assumption of continuity of the spin vector field [74]. The technique has also
recently been applied to the imaging of 2D materials [75], magnons [76] and
antiferromagnets [77].

Spin-Polarized Scanning Tunneling Microscopy

In scanning tunneling microscopy (STM), the apex of a conductive tip is placed


near the surface of a conductive sample. A bias voltage is applied between sample
and tip, and a small tunneling current flows that decays exponentially with the tip-
sample separation. In the constant current mode, a feedback mechanism adjusts the
tip-sample distance such that the tunneling current is kept constant. When the tip is
scanned over the surface, the tip apex moves on lines of constant current, which are
related to first order to lines of constant density of states, i.e., reflecting the sample
topography. For spin-polarized STM, a spin-polarized tunneling current is needed
that in principle can be obtained in various ways [78, 79].
Before the already mentioned use of ferromagnetic tips is discussed, we shortly
report on alternative approaches that have been tried with less success. The possi-
bility to photo-excite spin-polarized carriers from GaAs tips has been considered
by Suzuki et al. [80]. Circularly polarized light was used to pump spin-polarized
carriers into the conduction band of the tip that then tunnel into the sample. By
modulating the helicity, the tunneling current is modulated due to spin-dependent
tunneling, which can be detected by a lock-in amplifier. The signal can be used to
separate spin information from the sample topography. However, this method suffers
1220 R. M. Reeve et al.

from low contrast and additional magneto-optical contrast of low resolution. As a


new development, photoemission or photoassisted tunneling from metallic tips has
been proposed for the introduction of time resolution to STM on a femtosecond
level [81]. Using the inverse effect, Alvarado et al. [82] measured the circular
polarization of the light that is emitted because of the tunneling current. This method
suffers from the very low quantum efficiency of the inverse photoemission effect,
and details of the emission process are still debated today [83].
For ferromagnetic tips, the separation of topography and spin information is an
important issue. When a finite negative (positive) bias voltage, V , is applied to the
sample with respect to the tip, the occupied sample (tip) states in the range of width
eV below the Fermi level of the sample (tip) contribute to the tunneling. In the
tunneling process, the electrons tunnel into the unoccupied tip (sample) states of
the range eV above the Fermi level. The spin polarization of both the tip and the
sample states contribute to the tunneling. Therefore, the spin polarization of the
tunneling current varies with sample bias. Variations of the tunneling conductance
are compensated by the feedback loop and show up in the topography that then
contains both topographic and magnetic information.
An effective way of separating magnetic and topographic information has been
demonstrated by Wulfhekel et al. [57], who modulated the tip magnetization by
a small coil and detected the modulation of the tunneling current by a lock-
in amplifier. For out-of-plane sensitivity, the coil is wound around the tip axis.
For in-plane sensitivity, one uses a small ring from a soft magnetic material,
where the outer rim is used as a tip with surprisingly good resolution. The tip
material has to be chosen carefully as magnetostriction produces an additional
signal with the same frequency as the modulation. Domain wall widths in Mn/Fe
bilayers in the range of 1 nm and step-induced frustration of antiferromagnetic
order have been resolved in this manner [84]. The most convenient way of
separating magnetic and topographic signals is by comparing two measurements
obtained for opposite magnetization directions of tip or sample [85]. One option
is to repeat the measurement after an external field has rotated tip or sample
magnetization [86]. A second possibility is to compare sample areas of supposed
identical chemical structure but opposite magnetization, e.g., an epitaxially grown
nanowire of constant thickness [87]. A third option is to repeat the measurement
with intentionally changed tip magnetization, e.g., from out-of-plane to in-plane
magnetization [88].

Technical Details
Chemically etched tungsten tips are the most commonly used tips for STM. Starting
from these tungsten tips, a thin ferromagnetic film evaporated on the tip apex then
serves as a ferromagnetic counter electrode. The stray field of the thin film is small
enough to avoid dipolar sample-tip interaction. Alternatively, antiferromagnetic tips
are used. In this case, Cr tips etched from thin Cr wires are advantageous. For
SP-STM, the tip must be prepared in situ in UHV, for example, by voltage pulses,
24 Magnetic Imaging and Microscopy 1221

in order to obtain high resolution and magnetic contrast. Usually, only a single
magnetization component is detected. The sensitivity axis of the tip is in many cases
not obvious and must be calibrated on known magnetization structures.
The tunneling current can be described as a sum of a spin-averaged and a spin-
dependent term [89]. Following Ref. [90], the tunneling current can be calculated
by Fermi’s golden rule:
 ∞
I∝ |M(ε)|2 ρT (ε − eV )ρS (ε)[f (ε − eV ) − f (ε)]dε. (3)

Here, M denotes the tunneling matrix element; ρT ,S is the spin-dependent density


of states for tip and sample, respectively, and f is the Fermi function. As a result
of the integration, the spin-polarized contribution to the current becomes reduced
if the spin polarization changes sign between εF and eV . In order to increase the
magnetic contrast and also for separation of topographic and magnetic information,
it is helpful to measure the differential conductance dI /dV , which in the case of
low temperature and low bias is given by [78]

dI /dV (V ) ∝ |M(ε)|2 ρT ρS (eV − εF ). (4)

Here, we have assumed an energy-independent value for ρT (εF ). By introducing


the spin polarization for tip, PT (εF ), and sample PS (eV − εF ), the differential
conductivity can be written as

dI /dV (V ) = g0 (V )[1 + PT (εF )PS (eV − εF )cosφ], (5)

with φ denoting the angle between tip and sample magnetization and g0 the
unpolarized conductivity. The assumption of a constant tip polarization ρT (εF ) is
realistic in the case of V > 0 probing the unoccupied states of the sample because
the tunneling current is dominated by electrons from the Fermi level of the tip. For
illustration of this case, the tip and sample density of states is sketched on a common
energy scale in Fig. 6a. In contrast, when probing the occupied sample states V < 0,
a strong convolution with the tip density of states has to be considered.

Experimental Examples
The imaging of molecular structures is an important step toward the understanding
of spin transport and scattering in hybrid organic-metallic interfaces [91]. An
example of a Cu-phthalocyanine molecule on a metallic ferromagnetic surface is
shown in Fig. 6 [92]. Figure 6c–h shows dI /dV spectra and the corresponding
spin asymmetry, defined as A = [D(↑↑) − D(↑↓)]/[D(↑↑) + D(↑↓)]. From the
spectroscopic results obtained by collecting the current at tip positions over the
molecule, one can observe the almost entire suppression of the peaks at −0.1 eV,
as compared to the clean Fe surface. In contrast, a new peak appears at +0.4 eV
1222 R. M. Reeve et al.

Fig. 6 (a) Sketch of the tip and sample density of states in a common energy scale assuming
positive sample bias. (b) Topographical image of a Cu-phthalocyanine molecule on Fe(110).
dI /dV (V )-spectra (c, e, g) and asymmetries (d, f, h) extracted from the indicated areas in
(b) on the Cu-phthalocyanine molecule deposited on Fe(110). dI /dV (V )-spectra are measured for
parallel (black) and antiparallel (red) orientation of tip and sample magnetization. Experimental
asymmetries (blue) are compared to calculated asymmetries (black lines). For comparison, the
asymmetry of clean Fe, (d) scaled by 0.5, are also shown in red lines in (f) and (h). (Reprinted
figure with permission from [T. Methfessel, S. Steil, N. Baadji, N. Grossmann, K. Koffler, S.
Sanvito, M. Aeschlimann, M. Cinchetti, H. J. Elmers, Phys. Rev. B 84 224403 (2011)] Copyright
(2011) by the American Physical Society. [92])

which is attributed to electronic states that originate from the hybridization of


the lowest unoccupied molecular orbital (LUMO) of the free molecule with the
substrate. By comparing the asymmetry of the clean Fe layer to that of the CuPc
molecules, two regions are distinguished, denoted as regions 1 and 2. In region 1,
the spectral features of the asymmetry are only little modified with the exception of
a global reduction of A by about 50%, which is explained by assuming that the Fe-
CuPc interface acts as a featureless scattering barrier. The pronounced deviations
in region 2 are explained by the presence of spin-polarized hybridized interface
states.
The observation of atomic-scale magnetic skyrmions in ultrathin magnetic
films also highlights the ultrahigh resolution of spin-polarized STM for magnetic
microscopy [93]. The nontrivial spin textures are topologically stable, particle-like
spin configurations that can be used as information carriers. Spin-polarized scanning
tunneling microscopy can not only be used for the imaging of skyrmions [94, 95]
(see Fig. 7) but also for writing and deleting individual skyrmions [96] by employing
the tunneling electrons.
24 Magnetic Imaging and Microscopy 1223

Fig. 7 Spin-polarized STM images obtained with a Cr tip on an ultrathin PdFe/Ir(111) film
measuring the in-plane magnetization component. Detailed view of an isolated skyrmion. With
decreasing external magnetic field, the skyrmion gradually develops an elliptical shape and finally
stretches into a spiral domain. (Reprinted figure with permission from [A. O. Leonov, T. L.
Monchesky, N. Romming, A. Kubetzka, A. N. Bogdanov and R. Wiesendanger, New J. Phys.
18 065003 (2016)] Copyright (2016) by the Institute of Physics. [95])

Magnetic Force Microscopy

MFM is a special operation mode of atomic force microscopy employing a magnetic


probe, which interacts with the magnetic stray fields of the sample [97]. Therefore,
this technique measures the stray field distribution rather than the magnetization
structure itself. Recent developments are focused on the quantitative analysis
of data, the improvement of resolution, and the application of external fields
during measurement [98]. The interpretation of images acquired by MFM requires
knowledge about the specific near-field magnetostatic interaction between probe
and sample. In addition, one has to consider the properties of suitable probes. More
details can be found in Refs. [98, 99, 100, 101].

Technical Details
For the measurement of the magnetic forces, almost exclusively the dynamic
mode is applied, where resonance frequency shifts of the oscillating cantilever are
measured either directly or indirectly by the amplitude variation for fixed excitation
frequency. The frequency of the oscillating cantilever is given by


1 dFm
f = f0 1 − (6)
c∗ dz
1224 R. M. Reeve et al.

with f0 being the resonance frequency without interaction, dF m


dz the gradient of
the magnetic force, and c∗ an experimental constant. The sign of the frequency
shift distinguishes between attractive and repulsive forces (Δf < 0 and Δf > 0,
respectively). The most common detection method uses the amplitude signal and
is referred to as amplitude modulation. The cantilever is driven slightly away from
resonance, where the slope of the amplitude-versus-frequency curve is large. Mea-
surement sensitivity has an inverse dependence on the Q-factor of the oscillating
system that describes the damping. However, a high Q-factor has the drawback of
an increased response time of the detection system. In this case, a suitable alternative
is the frequency modulation (FM) technique. The cantilever self-oscillates with
constant amplitude A0 , with tip-sample interactions shifting the actual cantilever
frequency f by Δf = f − f0 .
The standard MFM probes are etched silicon tips with magnetic coatings
consisting of 10–150 nm Co/Cr multilayer structures and an effective magnetic
moment of around 10−22 Vsm. However, a large variation of materials have been
applied. A large coercive field is favorable in order to avoid a change of the magnetic
configuration of the tip during scanning.
For the separation of topography and magnetic signal, a constant distance mode
is applied. This lift mode involves measuring the topography on each scan line in a
first scan and the magnetic information in a second scan of the same line. This height
data of the first scan is used to move the tip at a constant local distance above the
surface during the second (magnetic) scan line, during which the feedback is turned
off. At this larger distance, the topographic interaction has decreased to a level that it
does not overlay the magnetic interaction, which decreases with distance at a smaller
rate.
The force interaction can be avoided by using diamond NV centers for the
detection of the stray field as described above . A schematic of this method is shown
in Fig. 8.

Experimental Examples
An example of high-resolution imaging using MFM is its employment to image
bit-patterned media with perpendicular anisotropy, where a resolution of better
than 10 nm has been demonstrated by evaluating line profiles in the images [102].
Figure 9 shows as a second example an image of a magnetic skyrmion indicating
the high resolution obtained by this method.
A considerable increase in spatial resolution can be achieved by magnetic
exchange force microscopy. The general concept of magnetic exchange force
microscopy relies on the combination of the atomic resolution atomic force
microscopy with spin sensitivity by using as a force sensor a magnetic tip mounted
on the free end of a cantilever. During scanning in the x-y plane, Δf is kept constant
by adjusting the z position of the tip relative to the surface so that the recorded
topographic image represents the condition of a constant tip-sample interaction
force. Selecting a more negative Δf set point increases the attractive interaction;
that is, the tip-sample distance is reduced. This method permits atomic resolution
24 Magnetic Imaging and Microscopy 1225

Fig. 8 Schematic of the NV center working principle. The NV center at the tip of an AFM is
rasterscanned across a magnetic surface at constant height. The stray field of an inhomogeneous
magnetization structure splits the states with mz = ±1 in proportion to the local field component
along the NV axis. By continuous illumination with green light through confocal optics, the mz = 0
state is excited and initiates red fluorescent light, detected by the same optical setup. If the ground
state is depopulated by resonant pumping with ca. 3 GHz microwave radiation into the mz = ±1
states, the fluorescent intensity will decrease. The resonance frequency is a measure for the stray
field component

on conducting and nonconducting surfaces in the noncontact regime with height


differences (or contrast) in the topography image reflecting variations of the short-
range forces. A purely chemical and structural contrast would reflect only the
arrangement of atoms. If a magnetic exchange interaction between tip and sample
is present, an additional contrast modulation occurs between neighboring rows of
magnetic atoms in an otherwise identical chemical environment. For this reason,
the exchange interaction can be distinguished unambiguously from other tip-sample
interactions. For the illustration of the method, we show a result of Kaiser et al. [58]
obtained for the surface of the antiferromagnet NiO. Figure 10 shows two atomically
resolved images for NiO(001). Both images were acquired on the same sample area.
The topographic image (Fig. 10a) recorded at a smaller frequency shift exhibits the
(1 × 1) symmetry of the chemical surface unit cell. In the Fourier transform of
the data, the chemical unit cell is represented by four spots. Figure 10b acquired
at a larger frequency shift, i.e., at smaller sample tip distance, shows an additional
modulation: every second row of nickel atoms along the [110] direction seems more
depressed, as indicated by the black arrows. The corresponding Fourier transform
(Fig. 10d) exhibits the appearance of one additional pair of peaks located halfway
between the center and two (opposing) peaks corresponding to the chemical unit
cell. This additional contrast modulation on neighboring nickel rows reflects the
antiferromagnetic surface unit cell of NiO(001).
1226 R. M. Reeve et al.

Fig. 9 High-resolution MFM image of a skyrmion at room temperature in an Ir-/Co-/Pt-based


multilayer. The out-of-plane stray field component is imaged. (Reprinted by permission from
Springer Nature: Nature Scientific Reports (How to measure the local Dzyaloshinskii-Moriya
Interaction in skyrmion thin-Film Multilayers, M. Baćani, M. A. Marioni, J. Schwenk and H. J.
Hug), Copyright (2019) [103])

Fig. 10 MFM images using the same tip at 7.9 K in 5 T on the same area. (a) Image recorded
at a constant frequency shift of −22.0 Hz. (b) Image measured at a constant frequency shift of
−23.4 Hz corresponding to a reduction of tip-sample distance of 30 pm. (c, d) Fourier transforms
of (a) and (b), respectively. (Reprinted by permission from Springer Nature: Nature (Magnetic
exchange force microscopy with atomic resolution, U. Kaiser, A. Schwarz and R. Wiesendanger),
Copyright (2007) [58])
24 Magnetic Imaging and Microscopy 1227

To summarize, spin-polarized STM offers subatomic resolution and spectro-


scopic information, which is certainly an advantage compared to MFM. This comes
at a cost: The surface sensitivity of STM requires a clean sample surface, which is
often difficult to maintain. The lateral resolution of MFM is limited by the inherent
tip-sample interaction. The NV center principle avoids this interaction, but the limits
of the spatial resolution have yet to be demonstrated. The advantage of the two latter
methods is that they can be applied under ambient conditions.

X-Ray Imaging

X-ray imaging is often the method of choice when studying magnetic materials in
operando conditions, primarily because x-ray absorption and scattering are largely
insensitive to the presence of other DC or AC electromagnetic fields. The resolution
is one to two orders of magnitude better than optical microscopy (see Table 1),
which makes x-ray imaging a powerful tool to study the texture and dynamics of
nanometer-scale magnetic spin structures. Moreover, compared to electron beams,
(sub-)ps x-ray pulses are often available in higher intensity and coherence. Finally,
the larger penetration depth of x-rays compared to electrons allows for imaging
of much thicker samples as compared to transmission electron microscopy. These
advantages typically come at the price of lower spatial resolution compared to
electron microscopy as well as a limited availability of laboratory-based sources.
In this section, we review the working principle as well as the strengths
and weaknesses of the most established x-ray imaging techniques. Specifically,
we discuss transmission x-ray microscopy (TXM), scanning transmission x-ray
microscopy (STXM), photoemission electron microscopy (PEEM), and coherent
x-ray imaging (CXI). Typical images of these techniques are shown in Fig. 11. In
the end, we also briefly present a technique for band structure or momentum space
imaging of magnetic materials, namely, spin-polarized angle-resolved photoemis-
sion spectroscopy (SP-ARPES). Reviews of imaging with x-ray microscopy can be
found in Refs. [104, 105].

X-Ray Magnetic Circular Dichroism: A Contrast Mechanism

X-ray magnetic circular dichroism (XMCD) [110] is the contrast mechanism


for almost all real-space x-ray magnetic imaging techniques. The XMCD effect
describes how the absorption of photons at a specific energy depends on the relative
orientation of the local magnetization and the helicity of the photons. XMCD
provides strong contrast to the element-specific magnetization component along the
beam direction (often the out-of-plane direction), which is a key distinction to the in-
plane contrast obtained by many electron microscopy techniques. Here, we briefly
review the excellent derivation of XMCD by Stöhr and Siegmann [111].
Consider the L-edge resonant photon absorption of a 3d magnetic transition
metal, i.e., the excitation from the localized 2p level to the delocalized and spin-
1228 R. M. Reeve et al.

Table 1 Comparison of magnetic imaging techniques, presenting some of the key specifications
and attributes. The quoted values are in general typical achievable values.∗ Proof of concept
recently demonstrated [48, 49, 152]
Probed Spatial Temporal
Technique quantity resolution resolution Info. depth Comments
Lorentz Stray field 10 nm 1 ns Sample Thin samples,
microscopy + average quantitative info. with
sample differential phase
induction contrast microscopy
Electron Stray field 5 nm 10 ms Sample Quantitative info.
holography + average through mathematical
sample image reconstruction
induction
SEMPA Magnetization 20 nm 700 ps∗ 1 nm Quantitative info.,
long acquisitions,
UHV required
SP-STM Magnetization Atomic 120 ps∗ Surface UHV required,
usually low
temperature,
long acquisitions
MFM Stray field 10–100 nm Low 1000 nm Potentially invasive,
long acquisitions,
few sample
requirements
NV mag- Stray field 10–20 nm Low 1000 nm Noninvasive,
netometry single spin sensitivity,
long acquisitions
TXM Magnetization 25 nm 50 ps Sample Synchrotron
average technique,
quick overview
images
STXM Magnetization 15–25 nm 50 ps Sample Synchrotron
average technique,
high repetition rates
PEEM Magnetization 40 nm 50 ps 5 nm Synchrotron
technique,
discharges possible
due to high potential
CXI Magnetization 10–30 nm fs-ps Sample Zero drift,
average synchrotron
technique,
complex sample
fabrication & image
reconstruction
MRI Proton density 1–2 mm 100 ms- 3D Low risk,
& several imaging very versatile
environment sec.
MEG Stray field 5 mm <1 ms 3D No unique solution,
imaging risk free
via
modelling
24 Magnetic Imaging and Microscopy 1229

Fig. 11 Typical images of magnetic skyrmions recorded by various x-ray imaging techniques.
(a: Reprinted by permission from Springer Nature: Nature Scientific Reports (Room-temperature
chiral magnetic skyrmions in ultrathin magnetic nanostructures, O. Boulle, J. Vogel, H. Yang,
S. Pizzini, D. de Souza Chaves, A. Locatelli, T. O. Menteş, A. Sala, L. D. Buda-Prejbeanu, O.
Klein, M. Belmeguenai, Y. Roussigné, A. Stashkevich, S. M. Chérif, L. Aballe, M. Foerster,
M. Chshiev, S. Auffret, I. M. Miron & G. Gaudin), Copyright (2016) [106]. b: Reprinted by
permission from Springer Nature: Nature Nanotechnology (Discrete Hall resistivity contribution
from Néel skyrmions in multilayer nanodiscs, K. Zeissler, S. Finizio, K. Shahbazi, J. Massey, F.
Al Ma’Mari, D. M. Bracher, A. Kleibert, M. C. Rosamond, E. H. Linfield, T. A. Moore, J. Raabe,
G. Burnell & C. H. Marrows), Copyright (2018) [107]. c: Reprinted by permission from Springer
Nature: Nature Communications (Spin-orbit torque-driven skyrmion dynamics revealed by time-
resolved X-ray microscopy, S. Woo, K. M. Song, H.-S. Han, M.-S. Jung, M.-Y. Im, K.-S. Lee,
K. Soo Song, P. Fischer, J.-I. Hong, J. W. Choi, B.-C. Min, H. C. Koo & J. Chang ), Copyright
(2017) [108]. d: Reprinted by permission from Springer Nature: Nature Nanotechnology (Field-
free deterministic ultrafast creation of magnetic skyrmions by spin-orbit torques, F. Büttner, I.
Lemesh, M. Schneider, B. Pfau, C. M. Günther, P. Hessing, J. Geilhufe, L. Caretta, D. Engel, B.
Krüger, J. Viefhaus, S. Eisebitt & G. S. D. Beach), Copyright (2017) [109].)

polarized 3d band, as depicted in Fig. 12. This transition is well described as a


first-order dipole transition. The selection rules for such dipole interactions require
Δ = ±1, Δs = 0, Δml = q, and Δms = 0, where  is the magnitude of the
orbital angular momentum, ml is the magnetic quantum number corresponding to
the orbital angular momentum, s = 12 is the spin angular momentum, ms is the spin
orientation, and q is the orbital angular momentum of the incident photon, which is
1 for right circularly polarized photons and −1 for left circularly polarized photons
1230 R. M. Reeve et al.

Fig. 12 Left: Schematic illustration of the XMCD effect. X-ray illumination of the sample excites
element-specific electronic transitions from core levels to empty states at the Fermi level. Due
to the spin-split density of states in the magnetic specimen, the x-ray absorption cross sections
are different for negative and positive helicity of the incident photons, depending on the relative
alignment of the incident wave vector and the magnetization in the sample, as shown on the right
in the case of the L2 and L3 absorption edges of Fe. (Reprinted by permission of Springer Nature:
Springer (Magnetism: From Fundamentals to Nanoscale Dynamics, Interactions of polarized
photons with matter, p. 390, J. Stöhr and H. C. Siegmann), Copyright (2016) [111])

(all in units of h̄). In particular, we will use the fact that ml changes by q and that
ms is conserved.
The 2p levels experience a strong spin orbit coupling, typically on the order
of 15 eV, which ensures that the L2 = 2p1/2 → 3d and the L3 = 2p3/2 → 3d
absorption edges do not overlap. Consider the L2 transition, schematically illus-
trated in Fig. 13. The absorption intensity is proportional to the matrix element
I ∝ i,f | f |Cq |i | , where |i = |n = 2, l = 1, j = 1/2, mj = ±1/2 :=
1 2

|2p1/2 , mj is one of the two possible initial states, |f = |n = 3, L = 2, mL =


{−2..., 2}, ms =↑ := |3d ↑, mL is one of the five possible final states in the spin-
up branch, and Cq1 is the Racah spherical tensor that describes the angular part of
the dipole transition operator corresponding to an absorbed photon of helicity q.
Now consider the excitation of a |2p1/2 , mj = −1/2 state, which can be written
as a mixture of 2/3 of |ml = −1, ↑ and 1/3 of |ml = 0, ↓ . The coefficients are
the so-called Clebsch-Gordan coefficients, of which we here provide the squared
magnitudes because the absorption cross section is proportional to the squared
magnitude of the spin projection, e.g., | ↑ |2p1/2 , mj = −1/2 |2 = 2/3. The
spin-down contribution of the initial state cannot be excited into the spin-up d-band
since the optical excitation preserves the spin quantum number. Therefore, also the
orbital angular momentum is fixed, here to ml = −1. For each value of q, there is
exactly one possible final state, as illustrated in Fig. 13. In this case of the 2p to 3d
transition, the matrix element | f |Cq1 |i |2 is proportional to (3 + qml )(2 + qml ),
resulting in the transition rates indicated in Fig. 13. Multiplying the transition rates
24 Magnetic Imaging and Microscopy 1231

Fig. 13 Illustration of how to derive the relative absorption difference of positive (q = +1) and
negative (q = −1) helicity x-rays at the L2 resonance assuming that only spin-up states are
available in the 3d band. For details, see text

with the spin-up component of the initial states results in the relative absorption
intensities of 75% for negative helicity light and 25% for positive helicity light at
the L2 edge. The intensities into a spin-down 3d band are inverted with respect to q.
In general, both spin-up and spin-down states are available in the valence
band. The total absorption cross section is then proportional to the spin-dependent
absorption intensities times the spin-dependent density of available final states.
Therefore, the absorption of right circular light minus the absorption of left circular
light measures the difference of available up states minus the density of available
down states in the 3d band, which is proportional to the magnetization of the sample
along the photon propagation direction.
Imaging with XMCD contrast requires highly monochromatic, circularly polar-
ized x-rays at suitable resonant energies. Typically, these are the L (800 [∼]eV)
or M (60 [∼]eV) edges of the magnetic transition metals, a range known as
the soft x-ray or extreme UV regime, respectively. Such light is available with
high intensity at modern synchrotrons and free electron lasers (FELs), and up to
now, most x-ray magnetic imaging is performed at these facilities. However, the
development of high harmonic generation sources has made tremendous progress
recently [112, 113], and laboratory-based XUV imaging of magnetic domains has
recently been demonstrated [114].
1232 R. M. Reeve et al.

TXM: Quick Full-Field Imaging in Transmission Geometry

Transmission x-ray microscopy (TXM) [115,116] can be seen as an analog to visible


light microscopy, with enhanced resolution by using smaller probing wavelengths.
Lenses at the wavelengths of soft x-rays are realized by diffractive elements, so-
called zone plates. The far-field diffraction pattern of a specimen is given by the
Fourier transform of its transmission function. A focus, i.e., a point-like diffraction
pattern, can be obtained from a Bessel function transmission function. A zone plate
is a binary version of a Bessel function absorption mask. Ultimately, the focus size
of a zone plate is determined by the width of the outermost zone. High-resolution
zone plates are difficult to fabricate and therefore very expensive.
The concept of TXM is illustrated in Fig. 14. Similar to an optical microscope,
TXM employs a condenser (KZP) that reduces the spot size of the incoming light to
the field of view of the subsequent objective lens, i.e., to a circle of approximately
10 µm in diameter. Like every zone plate, the condenser has a limited efficiency
on the order of 10 %. The majority of the transmitted light is undiffracted zero-
order light, which is blocked by an order selecting aperture (OSA). The focus of
the condenser depends on the wavelength of the incoming light. The position of the
OSA is optimized for transmitting the first-order light of the required wavelength,
blocking all other wavelengths because of their different cone angles. Hence, the
OSA also acts as a monochromator.
The light transmitted through the sample is collected by an objective zone plate
lens (MZP) and transformed to a real-space image of the local transmission intensity
of the sample on a CCD camera. Typically, the CCD camera has 2048 pixels per line
and is operated in 2 × 2 binning mode, resulting in 10 nm pixel size for the 10 µm

CCD
MZP

Sample
OSA

KZP

Fig. 14 Schematic illustration of a transmission x-ray microscope. The incident beam is transmit-
ted through a condenser zone plate (KZP), which is made of alternating opaque and transparent
rings to mimic a Bessel transmission function. The non-diffracted zero-order light, as well as
higher-order diffractions, is blocked using an order selecting aperture (OSA). The sample is placed
close to the focus of the KZP. An image of the transmitted light is generated via an objective zone
plate (MZP) on a CCD camera chip
24 Magnetic Imaging and Microscopy 1233

field of view. The binning allows for low-noise readout in 1 s, and a good quality
image is obtained by accumulating ∼20 images per helicity, yielding a full XMCD
image in approximately 1 min.
The fast acquisition of large-scale images with good resolution is the major
advantage of TXM compared to other magnetic imaging techniques. Furthermore,
some TXM end stations are built such that the OSA and the MZP act as vacuum
windows and the sample is in air, which is helpful for some applications. However,
due to the full-field nature of the technique, fast multi-pixel detection is needed for
dynamic imaging, limiting the repetition rate for dynamic processes.

STXM: Optimized for Dynamic Imaging

Scanning transmission x-ray microscopy (STXM) [116,117] is similar to TXM, but


instead of collecting a full-field image with an objective zone plate, the sample is
scanned with high precision through a focused x-ray spot, and the total transmission
is detected by a fast avalanche photodiode (APD). The resolution is now determined
by the spot size of the incident photons, which is typically 25 nm but can be
significantly smaller with more sophisticated zone plates (usually at the loss of total
intensity) or using ptychography [118,119,120] (at the price of readout speed). The
readout of the APD is extremely fast, indeed faster than the temporal separation,
δt, of subsequent x-ray flashes (typically 2 ns at synchrotrons). A configurable
number of channels, n, is available for counting the transmitted photons. Provided
the investigated sample is excited with an excitation of period δt n/m (with m, n
coprime integers, m often denoted “magic number”), a movie of the response of the
sample is directly obtained from the images collected in the n different channels.
The frames of such a movie can be reshuffled in a pulse-chronological order,
such that the dynamic response to the excitation is scanned in n temporal steps
of δt/m. Ideally, the number of channels is coprime to the number of bunches in the
synchrotron ring, such that the light from every bunch contributes equally to each
channel and bunch fluctuations are averaged out.
STXM is a very versatile technique. It allows for almost arbitrary zooming and
real-space translation of the field of view. That is, a large number of objects can be
investigated without changing samples. The capability of recording movies directly
makes the technique favorable for dynamic imaging. Typically, a single XMCD
image of a 3 µm field of view can be obtained in less than 2 min, and a full movie
with hundreds of frames can take less than 30 min. The image quality is mostly
determined by sample drift within the acquisition time, which is often visible in line
scan artifacts (see Fig. 11b).

PEEM: Imaging Surfaces of Bulk Samples

Photoemission electron microscopy (PEEM) makes use of the fact that an electron
excited by photon absorption can relax by transferring its energy to other electrons
at the Fermi energy [121, 122]. If this release of energy to the Fermi level happens
1234 R. M. Reeve et al.

detector

electrons

X-rays

20kV

Fig. 15 Schematic illustration of PEEM. The incident x-rays hit the sample under grazing
incidence, hence being absorbed in the surface-near region. The larger the absorption cross section,
the more energy is deposited near the surface, where secondary electrons are generated that can
leave the sample into the vacuum. The free electrons are accelerated by a strong electric field of
20 kV toward a detection column, where they are energy filtered and focused to a 2D detector. For
simplicity, the electron optics are not drawn in the schematic

near the surface of the sample, some electrons receive enough energy to leave the
sample into the vacuum. These free electrons can be accelerated by a high voltage
and focused by a series of electromagnetic lenses to a pixel detector (see Fig. 15).
The number of electrons at a specific pixel of the detector is proportional to the
number of electrons emitted at the corresponding position of the sample, which
in turn is proportional to the absorption of photons in the surface-near region.
Making use of the XMCD effect, an image of the magnetization in the direction
of the incident x-rays can be obtained. Due to the low mean free path of secondary
electrons inside the material, the sensitivity is restricted to the first few nm below
the surface. To increase the absorption efficiency in this surface-near region, PEEM
is typically operated with the x-rays hitting the sample at grazing incidence. PEEM
can also be realized in a scanning configuration [116]. Time-resolved measurements
are possible by synchronizing an external excitation of the sample with the x-ray
imaging pulses, thus allowing for pump-probe type measurements. In contrast to
the other x-ray imaging techniques discussed here, PEEM does not require an x-ray
transparent sample. Therefore, bulk substrates of arbitrary thickness can be used,
which makes PEEM very attractive to investigate epitaxial samples. A constraint is
that the sample surface must be conducting. PEEM is also a favorable technique for
imaging the in-plane components of the magnetization due to the grazing incidence
geometry [106], see Fig. 11a.
One example of PEEM imaging is presented in Fig. 16 where pump-probe
techniques have been employed to reveal the dynamic response of a magnetic
domain wall to a displacing field pulse [123]. A transverse domain wall is placed
in the center of a 2 μm wide half-ring by a static external field, as shown in (a).
Repeated pulsed in-plane fields are then generated to displace the domain wall by
24 Magnetic Imaging and Microscopy 1235

Fig. 16 Dynamic pump-probe PEEM imaging of the field-induced displacement of a transverse


domain wall. (a) and (b) show snapshots of the domain wall profile at different time delays,
revealing the wall motion. (c) By comparing the displacement field pulse (black) to the wall
displacement (green), the delayed onset of the wall motion is observed, indicative of domain wall
inertia. Furthermore, by subtracting the running average from the displacement, damped oscillatory
motion is revealed (data in blue, fit in red). (Reprinted figure with permission from [J. Rhensius,
L. Heyne, S. Krzky, L. J. Heyderman, L. Joly, F. Nolting, M. Kläui, Phys. Rev. Lett. 104 067201
(2010)] Copyright (2010) by the American Physical Society [123])

using 15 ps laser pulses, at a repetition rate of 63 MHz, which are synchronized


with the 70 ps x-ray imaging pulses from the synchrotron. The laser pulses generate
an electrical current pulse via a photodiode, and this in turn is passed through a
stripline to generate short, fast rise time field pulses as shown by the black trace
in (c). Pump-probe measurements with varying delay times then provide access
to the dynamic motion of the domain wall following the field pulse. A snapshot
of the domain wall displacement 200 ps following the onset of the field pulse is
shown in (b), and the full-time evolution of the domain wall displacement for a line
scan through the center of the wall is seen in (c). The onset of the wall motion is
observed to be delayed with respect to the field pulse, and then the wall is observed
to undergo damped oscillations (see inset), indicative of domain wall inertia, which
can be explained due to observed domain wall spin-structure changes before the
motion begins which act as an energy reservoir due to the increase of exchange
energy. By fitting the data, a domain wall oscillation frequency of 1.3 ± 0.6 GHz
has been extracted and a corresponding domain wall mass of (1.3 ± 0.1)×10−24 kg
deduced.

CXI: Zero Drift and Femtosecond Temporal Resolution

Coherent x-ray imaging (CXI) [124] is a common term for lensless imaging
techniques. Generally, the sample is illuminated with a coherent photon beam, and
the far-field scattering pattern, i.e., the squared magnitude of the Fourier transform
1236 R. M. Reeve et al.

of the sample’s transmission function, is recorded using a camera. The phase


information of the scattered wave is lost upon detection, and the reconstruction of
the transmission function of the sample is nontrivial. There are multiple approaches
to reconstruct the phase information. For example, known characteristic features of
the transmission function, such as an artificial binary mask before the sample, can
be used to computationally reconstruct the phase through iterative phase retrieval
algorithms. Phase retrieval is particularly popular in hard x-ray imaging and single
shot imaging of three-dimensional structures [124, 125].
The major drawback of phase retrieval imaging is the highly sophisticated image
reconstruction process that is required. Reliable and accurate reconstruction is
computationally expensive and requires deep knowledge of the algorithms and their
pitfalls. Recently, the concept of ptychography has been adapted to improve the
robustness of CDI phase retrieval by recording scattering patterns from multiple
largely, but not completely, overlapping regions of the specimen [126, 127].
However, up to now, phase retrieval and even ptychography are often too time-
consuming and involved to be competitive with other soft x-ray techniques when
it comes to magnetic imaging [118, 119, 120]. Ptychography is in this field mainly
used for ultrahigh-resolution imaging. For most applications aiming for 20 to 50 nm
spatial resolution, Fourier transform holography (FTH) [128, 129] has become the
CXI method of choice of magnetic imaging. The concept is illustrated in Fig. 17.
Essentially, the phase problem is solved by interference of the scattered beam with a
reference beam from a point-like reference source. If the reference source is laterally
separated from the specimen by a vector, r 0 , then the reference transmission
function is given by δ(r − r 0 ). Applying a Fourier transform to the scattering
interference pattern, the so-called hologram, yields the autocorrelation of the total
transmission function. This autocorrelation includes the cross correlations between
the specimen and the reference delta function, hence reconstructing the specimen’s
transmission function without any sophisticated algorithm. Figure 17 shows an
example of a sample used to image magnetic skyrmions and a reconstruction of the
magnetic pattern obtained from an inverse Fourier transform of the hologram. Note
that CXI can simultaneously reconstruct the amplitude and phase of the transmitted
exit wave [130, 114], similar to electron holography. However, in the magnetic
materials studied so far with CXI, amplitude and phase were found to encode the
same magnetic information [131], namely, the magnetization along the beam, which
is why magnetic CXI images are typically presented using a one-dimensional color
scale.
CXI has its strengths in drift-free imaging and in single shot destructive imaging
at free electron laser (FEL) sources [132, 133]. Drift-free imaging can be realized if
one ensures that the scattering always originates from the same area of the sample,
either due to a mask that is monolithically integrated with the sample or because
the sample size is so small that the entire sample is imaged at once. The remaining
drift between sample and detector is much less of concern since it only matters on
a micrometer length scale – the size of a camera pixel – which is relatively easy
to suppress mechanically or by centering each scattering pattern, and even if left
uncorrected only results in a phase shift after Fourier-transforming the scattering
24 Magnetic Imaging and Microscopy 1237

Fig. 17 Schematic illustration of x-ray holography. The main image illustrates the principle of
x-ray holography, where a coherent x-ray beam (blue) illuminates a specimen behind a circular
aperture and a close-by reference hole. The scattered beams of these two objects interfere on
a CCD camera chip, forming the hologram. The transmission function of the sample can be
reconstructed from the hologram via an inverse Fourier transform. Within the field of view, the
reconstruction shows two magnetic skyrmions (white) on a black background. Inset: Scanning
electron micrograph of a typical sample. The horizontal wire in the center is a magnetic multilayer.
The large features at the left and right edges of the image are gold contact pads for dynamic
experiments. The bright vertical rectangle is a SiN membrane. On the back side of the membrane
is a 1.5 µm thick gold layer with three holes in it, one 800 nm diameter hole defining the field of
view of the imaging (visible as a black shadow in the center of the image) and two smaller holes
defining the reference beam (visible as two circles below the magnetic wire). The white scale bar
is 5 μm long

pattern to obtain a real-space image. It has been demonstrated that the dynamics of
skyrmions can be tracked with 3 nm precision due to this intrinsic stability [134].
Furthermore, lensless imaging is so far the only viable technique for single shot
imaging at FELs, where the beam intensity is so high that all optical elements would
be destroyed. However, these advantages come at the price of a sophisticated sample
fabrication process and, up to now, a lack of permanent user facility end stations
(with some being under construction at NSLS-II, at PETRA-III, and at MAX-IV).

SP-ARPES: Microscopy in Momentum Space

Beyond real-space imaging of magnetic domains, the local spin-dependent


band structure is of high relevance. Angle-resolved photoemission spectroscopy
(ARPES) with spin detection is capable of providing such information [135].
ARPES analyzes the electrons released from the sample surface upon photon
irradiation. However, instead of secondary electrons (as in the case of PEEM),
electrons emitted by direct transitions are detected. In this case, energy and
momentum conservation apply. Thus, initial state properties, i.e., binding energy,
momentum, and spin, can be deduced from the measured spectra. Recently, a
1238 R. M. Reeve et al.

time-of-flight momentum microscope [136] has been developed that is capable of


parallel detection of momentum, energy, and spin. The electron-optical setup is
related to PEEM, but instead of the sample surface, the backfocal (or diffraction)
plane is imaged on the detector. For more details, the reader is referred to the
specialized literature [137, 136, 37, 138].

Medical Magnetic Imaging

A number of magnetism-based imaging techniques exist which are employed in


a medical context [139]. In this section, we provide an overview of two of the
main modern magnetic-based techniques, magnetic resonance imaging (MRI) and
magnetoencephalography (MEG).

Magnetic Resonance Imaging

Magnetic resonance imaging employs the phenomenon of nuclear magnetic reso-


nance (NMR) in order to detect the high-field-induced polarization of the nuclear
spin state of selected species. Most commonly, the technique is set to be sensitive
to protons which are in high abundance in water and fat-rich regions of the
body. Depending on the precise sequence of magnetic field pulses applied during
the measurement, the technique provides a wide range of contrast mechanisms,
allowing for imaging of a whole range of different tissues with high spatial
resolution. Furthermore, MRI is noninvasive and does not expose the subject
to dangerous ionizing radiation, making it largely risk-free, although the large
magnetic fields involved exclude its use in patients with some implants. One of
the main challenges involved with the technique is the requirement to produce and
work with very uniform high magnetic fields which makes MRI machines large,
bulky pieces of equipment [140]. Superconducting coils are usually employed for
this purpose, which need to be kept at cryogenic temperatures. Care also needs to be
taken in the design and shielding of the required radio frequency (RF) coils for
detecting the signals [140]. Overall, the required technical infrastructure makes
this imaging rather expensive. Furthermore, typical scans can last tens of minutes,
which can be a challenge for the subject who is required to remain still during
this time. The constant cycling of gradient magnetic fields within the instrument
also leads to vibrations and associated loud noise, which is an additional source
of discomfort for the patient. Nevertheless, the flexibility of MRI and wealth of
possible information mean that the technique is widely used to image all parts of the
body for anatomical determinations such as detecting tumors and brain imaging. In
the following, we outline the main principles of the technique. Further details on the
wide range of imaging modes and their applications are provided in the dedicated
literature [141, 142, 143].
24 Magnetic Imaging and Microscopy 1239

Nuclear Magnetic Resonance

A nucleus with nonzero spin, such as hydrogen, in an external magnetic field along
the z axis, Bz , will experience a Zeeman splitting of the otherwise degenerate
energy levels corresponding to different z components of the nuclear angular
momentum [144]. This yields a spontaneous polarization of the system; however,
since the energy splitting is very low compared with typical thermal energies, the
resulting polarization is very small. In order to have an appreciable signal, very
strong magnetic fields are required, which tend to be of the order 0.3–2 T in the used
instrumentation, although they can be much higher. Manipulation of the resulting
magnetization is now possible through the application of suitable radio frequency
(RF) field pulses which need to be set around the resonant frequency of the
system for appreciable effects. This frequency is known as the Larmor frequency:
ωL = γ Bz where γ is the gyromagnetic ratio. The large natural abundance of 1 H
in a human body, in combination with its comparatively large gyromagnetic ratio,
make hydrogen an attractive choice for imaging biological systems; however, other
nuclei such as phosphorus or sodium are also sometimes chosen. Since the Larmor
frequency is also dependent on the chemical environment of the proton, a so-called
chemical shift, it is possible to gain contrast based on the unique fingerprint of par-
ticular molecules or on the general environment of the nucleus. In some cases, if no
suitable contrast can be obtained from the naturally occurring differences in tissue
within the body, contrast-enhancing agents may be administered, e.g., intravenously.
These are typically paramagnetic gadolinium compounds or superparamagnetic iron
oxide nanoparticles which modify the magnetic environment of the imaging region.
The NMR experiment is schematically depicted in Fig. 18. In the initialized state,
the magnetization is aligned with the strong external field, Bz . An RF field pulse is
then applied via a coil around the sample, which acts to excite the magnetization and
rotate it away from the +z axis to an extent which is determined by the intensity and
duration of the RF pulse, as shown in Fig. 18a. Following the removal of the pulse,
the magnetization will decay back to the initial state. The relaxation dynamics of
the magnetization back to its equilibrium value, M0 , are described by the Bloch
equations:

dMx Mx
= γ (M × B)x − ∗ , (7)
dt T2
dMy My
= γ (M × B)y − ∗ , (8)
dt T2
dMz Mz − M0
= γ (M × B)z − . (9)
dt T1

The first term in each equation describes the precession of the magnetization
around the Bz field at the characteristic Larmor frequency (ω = ωL ). This
precession of the magnetization leads to a changing magnetic flux which can be
detected as an induced voltage in a series of pickup coils outside of the imaged
1240 R. M. Reeve et al.

Fig. 18 Schematic depiction of the NMR experimental setup. The magnetization vector is
represented in the shown coordinate system and arises due to polarization of the investigated nuclei
in a strong Bz field. (a) RF pulses tuned to the Larmor frequency of the system, ωL , excite the
magnetization, and cause it to rotate away from the z axis to an extent depending on the amplitude
and length of the excitation. (b) When the magnetization is not aligned with the Bz field, it will
precess around the z axis at a frequency ω = ωL . The flux from the magnetization dynamics is
picked up by coils and amplified to generate the signal. (Reprinted by permission of Wiley: (D. P.
Plewes & W. Kucharczyk, Physics of MRI: A Primer, J. Magn. Reson. Imaging 35, 1038, 2012),
Copyright (2012) [142])

object, as shown schematically in Fig. 18b. The same coils can used for detecting
the signal as those employed for the generation of the RF fields that manipulate the
magnetization. The second term in the first two equations describes the relaxation
of the transverse magnetization over a characteristic timescale known as the T2∗
relaxation time. This has contributions due to spin dephasing both as a result of
inhomogeneities in the local effective field that are fixed in time due to the varying
environment of different spins and also dephasing due to slowly time-varying field
variations and spin-spin relaxation processes. These are sometimes referred to as the
T2 and T2 relaxation times, respectively, with T1∗ = T1 + T12 . The second term in the
2 2
last equation describes the relaxation of the longitudinal magnetization and involves
24 Magnetic Imaging and Microscopy 1241

the dissipation of the excitation energy of the system back to the lattice over a time
characterized by the spin-lattice relaxation time, T1 .
Measurements of the NMR precessional signal decay over time are sensitive
to these various relaxation processes. By changing the measurement scheme, the
relative contribution of the T1 and T2 times to the relaxation can be varied, and
since these values are also influenced by local environments, they provide a flexible
range of contrast conditions for the resulting images, in addition to the basic
proton density contrast. T1 , for example, generally increases with the strength of
the applied field, whereas the T2 values are relatively insensitive to this and hence
field-cycling experiments where the strength of the uniform field is changed allow
one to investigate this dependence. Furthermore, by setting the duration of the
excitation pulse, the initial state can be changed which also yields different dynamic
contributions. For example, with a so-called π/2 pulse, the magnetization is nutated
into the x-y plane and will undergo both precession and relaxation. However, with
a π pulse, the magnetization rotates to the −z axis, and since it is colinear with the
field, no precession occurs and the decay is solely determined by the T1 time. If the
magnetization is subsequently rotated back to the x-y plane to generate a precession
signal, the initial amplitude provides contrast weighted by T1 .

Imaging and Pulse Sequences


In order to build up an image using NMR, it is necessary to encode the spatial
information about the region of signal generation in the global signal detected by
the pickup coils. This is achieved by using time-varying field gradients of the order
of 10 mT/m which are superimposed on the uniform global field. This provides
spatially varying Larmor frequencies, allowing one to selectively probe particular
spatial regions of the system with the choice of excitation frequency. A variety
of schemes exist to exploit this in imaging. For example, if a gradient field along
the z axis is applied during the initial excitation, only a particular slice will be in
resonance with the RF pulse, and hence, this slice is selectively probed in the mea-
surement. The thickness of the slice is determined by the gradient of the field and the
bandwidth of the excitation pulse. For spatial localization within the slice, x-y field
gradients are subsequently applied, and a Fourier approach is used to reconstruct
the real-space image. Due to the resulting different precession frequencies for the
different regions of the slice, the nuclear moments, which are initially in-phase,
begin to dephase, and hence the signal as a function of time progressively represents
different spatial frequencies of the image. By appropriately varying the pulses, field
gradients, and the time delay between excitation and measurement, the whole of
k-space can be probed, and by mathematical transformation, the real-space image
can be extracted.
Depending on the measurement scheme, a very wide variety of field pulse
sequences are applied to the sample which vary the weighting of the measurement
to the different relaxation times and vary the order in which k-space is sampled.
A typical sequence may subsequently apply incrementing field gradients in the z,
y, and x axes such that k-space is sampled in a Cartesian scheme. More advanced
schemes can be devised in order to map out radial or spiral k-space trajectories
1242 R. M. Reeve et al.

Fig. 19 The spin-echo approach. A 90◦ (π/2) pulse nutates the magnetization into the x-y plane
where it begins to precess. Dephasing occurs over a characteristic T2∗ time. A subsequent 180◦ (π )
pulse at time TE/2 reverses the order of the individual spins, which then begin to come together. At
a time TE, dephasing processes due to static field inhomogeneities characterized by T2 have been
corrected for. However, stochastic field fluctuations still play a role and contribute to the measured
signal. Additional decay of the signal will occur due to longitudinal magnetization relaxation as
characterized by T1 ; however, this is usually over longer timescales. (Reprinted by permission
of Wiley: (B. A. Jung & M. Weigel, Spin Echo Magnetic Resonance Imaging, J. Magn. Reson.
Imaging 37, 805, 2012), Copyright (2013) [145])

which can be more efficient, reducing the time required for data acquisition and
enabling imaging of dynamic processes [142]. Nevertheless, the process of scanning
a given part of the body can take quite a while. A further protocol can be employed
in order to counteract the dephasing caused by non-time-varying spatial gradients in
the global field and thereby separate this contribution to the T2∗ relaxation time. One
approach is known as a spin-echo technique [145], as indicated in Fig. 19. In the
first stage, a π/2 pulse is applied to nutate the magnetization into the x-y plane. The
system is then allowed to precess for a certain time, τ , during which the spins will
dephase due to the spatially inhomogeneous local fields. Next, a π pulse is applied,
after which the phase differences between the spins will have been reversed. Further
precession will gradually bring the spins back to their original in-phase state after
the echo time TE = 2τ . In this manner, the influence of field inhomogeneities and
the spread in fields due to chemical shifts are corrected for, and the measurement is
able to probe the spin-spin interactions from neighboring nuclei.
The spin-echo pulses are applied with a certain repetition time, TR, as the sample
is scanned. The length of TR in relation to the longer T1 relaxation processes, as well
24 Magnetic Imaging and Microscopy 1243

Fig. 20 MRI images of a


brain taken using a spin-echo
technique with different
values of the echo time, TE,
and repetition time, TR,
revealing different contrast in
each case. For short TR and
TE, the image is weighted by
the T1 time; for long TR and
TE, the image is weighted by
the T2 time; and for long TR
and short TE, the image
contrast reveals proton
density (PD). (Reprinted by
permission of Wiley: (B. A.
Jung & M. Weigel, Spin Echo
Magnetic Resonance
Imaging, J. Magn. Reson.
Imaging 37, 805, 2012),
Copyright (2013) [145])

as the choice of TE strongly, affects the weighting of the image contrast with respect
to the T1 and T2 relaxation times, as shown in the example in Fig. 20.

Functional Magnetic Resonance Imaging


Functional magnetic resonance imaging (fMRI) is one of the newer imaging modes
which is employed to investigate activity in the brain, in particular for fundamental
research studies. It relies on the change in the magnetic properties of blood cells
depending on their oxygenation state. Oxygen in the blood is carried by the protein
hemoglobin which is paramagnetic in its deoxygenated state. This paramagnetism
modifies the local field felt by nearby water molecules, thereby impacting the
effective T2 values in the vicinity of blood vessels carrying deoxygenated blood as
compared to those carrying oxygenated blood. This contrast mechanism is known
as blood-oxygen-level dependent or BOLD. This can be employed for functional
brain imaging since neural activity in a certain region is accompanied by a spike in
the delivery of oxygenated blood to that region.

Nuclear Quadrupole Resonance Imaging

Nuclear quadrupole resonance (NQR) is a related technique to NMR which can be


used for imaging in an analogous manner to MRI [146]. Many of the underlying
concepts of signal and image generation, as well as the required experimental setup,
are identical. The main difference is that the technique does not require the strong
1244 R. M. Reeve et al.

uniform background field to provide the energy splitting of nuclear states, but
rather the splitting is determined by the interaction of nuclear quadrupole electric
moments and the electric field gradients around the nucleus. Such field gradients
are determined by the precise chemical environment of the nucleus, and as such
the energy splitting and associated resonant frequencies can be used for chemical
fingerprinting. NQR is restricted to nuclei with a spin quantum number ≥1 for which
the quadrupole moment is nonzero, making it complementary to NMR, offering
alternative contrast origins when imaging a given system and new possibilities for
contrast agents. As with NMR, the system is excited using RF field pulses, and the
relaxation of the system is detected via the free induction decay signal generated in
pickup coils. Imaging is achieved in a similar manner to MRI by applying spatially
varying RF fields or magnetic fields [147]. Field cycling NQR systems can be
employed to improve the sensitivity of the measurements, for example, in the case
of low abundance of the investigated nuclei, in an approach combining NQR with
NMR. In a typical measurement, the field at the sample is alternated between a high-
field and low- or zero-field condition. This is most readily achieved by physically
moving the sample into and out of a field region. In the high-field environment, the
large nuclear splitting occurs, polarizing the nuclei. The sample is then transferred
to a low-field region and the RF pulse applied at the appropriate NQR frequency
to excite the state. Finally, the sample is transferred back to the strong field, and
an NMR measurement is carried out. In addition to the chemical environment, the
resonance lines are strongly affected by physical parameters of the system, and
hence the technique has also been employed for imaging temperature, stress, and
pressure in a sample [148].

Magnetoencephalography

Magnetoencephalography [149, 150, 151] (MEG) is an emerging brain imaging


technique which senses brain functionality by detecting the magnetic fields outside
of the brain generated due to the ion currents associated with neuronal activity.
However, due to the extremely weak signals which are of the order 10–100 fT,
particularly sensitive magnetometers are required to detect the generated fields, and
even then the detected signals necessarily correspond to the simultaneous firing of
thousands of neurons from a small volume. The principal magnetometer of choice
is a dc SQUID which is an incredibly sensitive device for magnetic field detection.
A typical MEG setup consists of an array of hundreds of SQUIDs arranged in a
grid so as to detect the field over the whole scalp. The SQUID loops can be wound
in different configurations in order to directly detect the field strength, or by using
either multiple stacked coils or variously twisted loops, the out-of-plane or in-plane
field gradients can also be detected. Due to the need to operate the superconducting
elements at either liquid helium or liquid nitrogen temperatures, the whole array sits
at the bottom of a cryostat, the base of which is concave so as to be better molded to
the head of the patient who sits underneath. The very small signals being measured
require that the whole apparatus needs to be housed in a magnetically shielded
24 Magnetic Imaging and Microscopy 1245

room, or else compensation procedures need to be applied in order to correct for


distant magnetic field sources which would otherwise swamp the signal of interest.
Furthermore, care needs to be taken during measurement to distinguish the brain
signal from artifacts from fields originating from elsewhere in the body such as
from the heart or eye regions.
MEG has a number of advantages in brain imaging studies. Since the technique is
completely noninvasive, it is a particularly safe tool. It has good spatial resolution,
being able to localize brain activity with millimeter precision, however not quite
as accurately as fMRI. Meanwhile, the temporal resolution of the technique is
very competitive, being able to detect changes on a sub-ms timescale, much faster
than fMRI. It can be seen as a complementary imaging technique to the related
electroencephalography (EEG) which detects the concurrently generated electric
fields, since both techniques are sensitive to differently oriented current dipoles
within the cranium. MEG also tends to have a better spatial resolution than EEG,
and furthermore, it is less sensitive to the conductivity variations from the detailed
structure of the head which distort the EEG signal. Particularly active areas of
application include studies of epilepsy and autism.

The Inverse Problem


Given a known current distribution and a knowledge of the details of the surrounding
medium such as the geometry, electric permittivity, and magnetic permeability, it
is a relatively straightforward exercise to calculate the resulting electromagnetic
field distributions from Maxwell’s equations. In an MEG measurement, the task
is to calculate the current distributions that were responsible for generating the
measured field. Unfortunately, it has been shown that there is no unique solution
to this so-called inverse problem, and hence, progress requires the development
of simplified models of the system which are suitably constrained to yield a
physically relevant solution. The simplest descriptions of the current distributions
approximate the current sources as current dipoles in an equivalent dipole model.
An equivalent current dipole represents a spatial average of the source currents
within a small volume of the brain, characterized by its position, orientation, and
strength. Depending on the complexity of the model, different numbers of equivalent
current dipoles can be assumed in order to fit the data. For a quasi-continuous
model, the brain is divided into a large number of discrete volume cells known
as voxels, and each voxel is allocated three orthogonal equivalent current dipoles.
The task is then to determine the strength of each equivalent current dipole in
order to recreate the observed field distribution. Whichever model is used, the
parameters are iteratively adjusted, and the resulting field is calculated in order to
minimize the deviation between the model and the experimental data. However,
depending on the model chosen, the problem can be severely underdetermined,
and in any case due to the lack of a unique solution, it is necessary to sensibly
constrain the problem. Models of the head usually approximate it as a uniform
spherical conductor. Improved modeling and constraints to the fitting algorithms
can be provided by complementary MRI imaging of the brain to provide accurate
anatomical details.
1246 R. M. Reeve et al.

Summary

As has been presented, a wide variety of techniques are available to image the
magnetic state of a system, and for a given application, it is necessary for the
user to judge the most appropriate option depending on the type of specimen,
the information that one wants to acquire, and the required spatial and temporal
resolution. To conclude this chapter, Table 1 provides a summary of some of the key
attributes of a selection of the most widely employed techniques to enable ease of
comparison. We note that the quoted values are not necessarily the ultimate limits
of the techniques, but rather in most cases represent typical values under standard
operating conditions.

Acknowledgments A large number of students, postdocs, colleagues, and collaborators have also
been involved in the authors’ research efforts on magnetic imaging over the years, only a few
examples of which we have been able to present here. Without these individuals, this work would
not have been possible, and we gratefully acknowledge all their contributions and insights.

References
1. Hubert, A., Rudolf, S.: Magnetic Domains: The Analysis of Magnetic Microstructures.
Springer, Berlin (2011)
2. Cowburn, R.P.: Property variation with shape in magnetic nanoelements. J. Phys. D: Appl.
Phys. 33, R1–R16 (1999)
3. Parkin, S.S.P., Hayashi, M., Thomas, L.: Magnetic domain-wall racetrack memory. Science
320(5873), 190–194 (2008)
4. Hrkac, G., Dean, J., Allwood, D.A.: Nanowire spintronics for storage class memories and
logic. Philos. Trans. R. Soc. A: Math. Phys. Eng. Sci. 369(1948), 3214–3228 (2011)
5. Diegel, M., Glathe, S., Mattheis, R., Scherzinger, M., Halder, E.: A new four bit magnetic
domain wall based multiturn counter. IEEE Trans. Magn. 45(10), 3792–3795 (2009)
6. Lee, J.-Y., Lee, K.-S., Choi, S., Guslienko, K.Y., Kim, S.-K.: Dynamic transformations of the
internal structure of a moving domain wall in magnetic nanostripes. Phys. Rev. B 76, 184408
(2007)
7. Boulle, O., Malinowski, G., Kläui, M.: Current-induced domain wall motion in nanoscale
ferromagnetic elements. Mater. Sci. Eng. R: Rep. 72(9), 159–187 (2011)
8. Bitter, F.: Experiments on the nature of ferromagnetism. Phys. Rev. 41, 507–515 (1932)
9. Hopster, H., Oepen, H.P.: Magnetic Microscopy of Nanostructures. Springer, Berlin/Heidel-
berg (2010)
10. Petford-Long, A.K.: Magnetic imaging. In: Ziese, M., Thornton, M.J. (eds.) Spin Electronics,
p. 316. Springer, Berlin/Heidelberg (2001)
11. Dahlberg, E.D., Proksch, R.: Magnetic microscopies: the new additions. J. Magn. Magn.
Mater. 200(1), 720–728 (1999)
12. Cheong, S.-W., Fiebig, M., Wu, W., Chapon, L., Kiryukhin, V.: Seeing is believing:
visualization of antiferromagnetic domains. npj Quantum Mater. 5, 3 (2020)
13. McCord, J.: Progress in magnetic domain observation by advanced magneto-optical
microscopy. J. Phys. D: Appl. Phy. 48, 333001 (2015)
14. Soldatov, I.V., Schäfer, R.: Advances in quantitative Kerr microscopy. Phys. Rev. B 95,
014426 (2017)
15. Soldatov, I.V., Schäfer, R.: Selective sensitivity in Kerr microscopy. Rev. Sci. Instrum. 88(7),
073701 (2017)
24 Magnetic Imaging and Microscopy 1247

16. Steierl, G., Liu, G., Iorgov, D., Kirschner, J.: Surface domain imaging in external magnetic
fields. Rev. Sci. Instrum. 73(12), 4264–4269 (2002)
17. Amelinckx, S.: Electron Microscopy: Principles and Fundamentals. VCH, Weinheim
(1997)
18. Helmut, K., Parkin, S.S.P.: Handbook of Magnetism and Advanced Magnetic Materials,
vol. 3. Wiley, Chichester (2007)
19. Hale, M.E., Fuller, H.W., Rubinstein, H.: Magnetic domain observations by electron
microscopy. J. Appl. Phys. 30(5), 789–791 (1959)
20. Chapman, J., Scheinfein, M.: Transmission electron microscopies of magnetic microstruc-
tures. J. Magn. Magn. Mater. 200(1), 729–740 (1999)
21. Zweck, J., Schneider, M., Sessner, M., Uhlig, T., Heumann, M.: Lorentz electron microscopic
observation of micromagnetic configurations in nanostructured materials. In: Kramer, B. (ed.)
Advances in Solid State Physics, vol. 41. Springer, Berlin/Heidelberg (2001)
22. Volkov, V., Zhu, Y.: Lorentz phase microscopy of magnetic materials. Ultramicroscopy 98(2),
271–281 (2004)
23. Chapman, J.N.: The investigation of magnetic domain structures in thin foils by electron
microscopy. J. Phys. D: Appl. Phys. 17, 623–647 (1984)
24. Chapman, J., Morrison, G.: Quantitative determination of magnetisation distributions in
domains and domain walls by scanning transmission electron microscopy. J. Magn. Magn.
Mater. 35(1), 254–260 (1983)
25. Dekkers, N., de Lang, H.: Differential phase contrast in a STEM. Optik 41, 452 (1974)
26. Krajnak, M., McGrouther, D., Maneuski, D., Shea, V.O., McVitie, S.: Pixelated detectors
and improved efficiency for magnetic imaging in STEM differential phase contrast. Ultrami-
croscopy 165, 42–50 (2016)
27. Mankos, M., Cowley, J.M., Scheinfein, M.R.: Quantitative micromagnetics at high spatial
resolution using far-out-of-focus STEM electron holography. Physica Status Solidi (A)
154(2), 469–504 (1996)
28. Cowley, J.: Twenty forms of electron holography. Ultramicroscopy 41(4), 335–348 (1992)
29. Midgley, P.A., Dunin-Borkowski, R.E.: Electron tomography and holography in materials
science. Nat. Mater. 8, 271–280 (2009)
30. Hawkes, P.W.: Aberration correction past and present. Philos. Trans. R. Soc. A: Math. Phys.
Eng. Sci. 367(1903), 3637–3664 (2009)
31. McVitie, S., McGrouther, D., McFadzean, S., MacLaren, D., O’Shea, K., Benitez, M.:
Aberration corrected Lorentz scanning transmission electron microscopy. Ultramicroscopy
152, 57–62 (2015)
32. Chrobok, G., Hofmann, M.: Electron spin polarization of secondary electrons ejected from
magnetized europium oxide. Phys. Lett. A 57(3), 257–258 (1976)
33. Jones, G.: Magnetic contrast in the scanning electron microscope: an appraisal of techniques
and their applications. J. Magn. Magn. Mater. 8(4), 263–285 (1978)
34. Kohashi, T., Konoto, M., Koike, K.: High-resolution spin-polarized scanning electron
microscopy (spin SEM). J. Electron Microsc. 59(1), 43–52 (2009)
35. Scheinfein, M.R., Unguris, J., Kelley, M.H., Pierce, D.T., Celotta, R.J.: Scanning electron
microscopy with polarization analysis (SEMPA). Rev. Sci. Instrum. 61(10), 2501–2527
(1990)
36. R. Allenspach Spin-polarized scanning electron microscopy. IBM J. Res. Dev. 44(4), 553–570
(2000)
37. Suga, S., Tusche, C.: Photoelectron spectroscopy in a wide hν region from 6ev to 8kev with
full momentum and spin resolution. J. Electron Spectrosc. Relat. Phenom. 200, 119–142
(2015)
38. Frömter, R., Hankemeier, S., Oepen, H.P., Kirschner, J.: Optimizing a low-energy electron
diffraction spin-polarization analyzer for imaging of magnetic surface structures. Rev. Sci.
Instrum. 82(3), 033704 (2011)
39. Burnett, G.C., Monroe, T.J., Dunning, F.B.: High-efficiency retarding-potential mott polar-
ization analyzer. Rev. Sci. Instrum. 65(6), 1893–1896 (1994)
1248 R. M. Reeve et al.

40. Reeve, R.M., Chin, S.-L., Ionescu, A., Barnes, C.H.W.: Modification of the secondary-
electron spin polarization in Co/Cu(110) films via gaseous adsorbates. Phys. Rev. B 84,
184431 (2011)
41. Lofink, F., Hankemeier, S., Frömter, R., Kirschner, J., Oepen, H.P.: Long-time stability of
a low-energy electron diffraction spin polarization analyzer for magnetic imaging. Rev. Sci.
Instrum. 83(2), 023708 (2012)
42. Abraham, D.L., Hopster, H.: Magnetic probing depth in spin-polarized secondary electron
spectroscopy. Phys. Rev. Lett. 58, 1352–1354 (1987)
43. Kuhrau, S., Kloodt-Twesten, F., Heyn, C., Oepen, H.P., Frömter, R.: Cap-layer-dependent
oxidation of ultrathin cobalt films and its effect on the magnetic contrast in scanning electron
microscopy with polarization analysis. Appl. Phys. Lett. 113(17), 172403 (2018)
44. VanZandt, T., Browning, R., Landolt, M.: Iron overlayer polarization enhancement technique
for spin-polarized electron microscopy. J. Appl. Phys. 69(3), 1564–1568 (1991)
45. Reeve, R.M., Mix, C., König, M., Foerster, M., Jakob, G., Kläui, M.: Magnetic domain
structure of La0.7Sr0.3MnO3 thin-films probed at variable temperature with scanning
electron microscopy with polarization analysis. Appl. Phys. Lett. 102(12), 122407 (2013)
46. Oepen, H., Benning, M., Ibach, H., Schneider, C., Kirschner, J.: Magnetic domain structure
in ultrathin cobalt films. J. Magn. Magn. Mater. 86(2), L137–L142 (1990)
47. Konoto, M., Kohashi, T., Koike, K., Arima, T., Kaneko, Y., Kimura, T., Tokura, Y.: Direct
imaging of temperature-dependent layered antiferromagnetism of a magnetic oxide. Phys.
Rev. Lett. 93, 107201 (2004)
48. Frömter, R., Kloodt, F., Rößler, S., Frauen, A., Staeck, P., Cavicchia, D.R., Bocklage, L.,
Röbisch, V., Quandt, E., Oepen, H.P.: Time-resolved scanning electron microscopy with
polarization analysis. Appl. Phys. Lett. 108(14), 142401 (2016)
49. Schönke, D., Oelsner, A., Krautscheid, P., Reeve, R.M., Kläui, M.: Development of a scanning
electron microscopy with polarization analysis system for magnetic imaging with ns time
resolution and phase-sensitive detection. Rev. Sci. Instrum. 89(8), 083703 (2018)
50. Schönke, D., Reeve, R.M., Stoll, H., Kläui, M.: Quantification of competing magnetic states
and switching pathways in curved nanowires by direct dynamic imaging. ACS Nano 14(10),
13324–13332 (2020)
51. Krautscheid, P., Reeve, R.M., Lauf, M., Krüger, B., Kläui, M.: Domain wall spin structures in
mesoscopic Fe rings probed by high resolution SEMPA. J. Phys. D: Appl. Phys. 49, 425004
(2016)
52. Koike, K., Kirschner, J.: Primary energy dependence of secondary electron polarization. J.
Phys. D: Appl. Phys. 25, 1139–1141 (1992)
53. Tersoff, J., Hamann, D.R.: Theory and application for the scanning tunneling microscope.
Phys. Rev. Lett. 50, 1998–2001 (1983)
54. Tersoff, J., Hamann, D.R.: Theory of the scanning tunneling microscope. Phys. Rev. B 31,
805–813 (1985)
55. Wiesendanger, R., Güntherodt, H.-J., Güntherodt, G., Gambino, R.J., Ruf, R.: Observation of
vacuum tunneling of spin-polarized electrons with the scanning tunneling microscope. Phys.
Rev. Lett. 65, 247–250 (1990)
56. Bode, M., Getzlaff, M., Wiesendanger, R.: Spin-polarized vacuum tunneling into the
exchange-split surface state of Gd(0001). Phys. Rev. Lett. 81, 4256–4259 (1998)
57. Wulfhekel, W., Kirschner, J.: Spin-polarized scanning tunneling microscopy on ferromagnets.
Appl. Phys. Lett. 75(13), 1944–1946 (1999)
58. Kaiser, U., Schwarz, A., Wiesendanger, R.: Magnetic exchange force microscopy with atomic
resolution. Nature 446, 522–525 (2007)
59. Shono, T., Hasegawa, T., Fukumura, T., Matsukura, F., Ohno, H.: Observation of magnetic
domain structure in a ferromagnetic semiconductor (Ga,Mn)As with a scanning Hall probe
microscope. Appl. Phys. Lett. 77(9), 1363–1365 (2000)
60. Chang, A.M., Hallen, H.D., Harriott, L., Hess, H.F., Kao, H.L., Kwo, J., Miller, R.E., Wolfe,
R., van der Ziel, J., Chang, T.Y.: Scanning Hall probe microscopy. Appl. Phys. Lett. 61(16),
1974–1976 (1992)
24 Magnetic Imaging and Microscopy 1249

61. Howells, G., Oral, A., Bending, S., Andrews, S., Squire, P., Rice, P., de Lozanne, A., Bland, J.,
Kaya, I., Henini, M.: Scanning Hall probe microscopy of ferromagnetic structures. J. Magn.
Magn. Mater. 196–197, 917–919 (1999)
62. Vu, L.N., Wistrom, M.S., Van Harlingen, D.J.: Imaging of magnetic vortices in superconduct-
ing networks and clusters by scanning squid microscopy. Appl. Phys. Lett. 63(12), 1693–1695
(1993)
63. Kirtley, J.R., Ketchen, M.B., Stawiasz, K.G., Sun, J.Z., Gallagher, W.J., Blanton, S.H., Wind,
S.J.: High-resolution scanning squid microscope. Appl. Phys. Lett. 66(9), 1138–1140 (1995)
64. Sidles, J.A., Garbini, J.L., Bruland, K.J., Rugar, D., Züger, O., Hoen, S., Yannoni, C.S.:
Magnetic resonance force microscopy. Rev. Mod. Phys. 67, 249–265 (1995)
65. Rugar, D., Budakian, R., Mamin, H.J., Chui, B.W.: Single spin detection by magnetic
resonance force microscopy. Nature 430(6997), 329–332 (2004)
66. Wagenaar, J.J.T., den Haan, A.M.J., de Voogd, J.M., Bossoni, L., de Jong, T.A., de Wit,
M., Bastiaans, K.M., Thoen, D.J., Endo, A., Klapwijk, T.M., Zaanen, J., Oosterkamp, T.H.:
Probing the nuclear spin-lattice relaxation time at the nanoscale. Phys. Rev. Appl. 6, 014007
(2016)
67. Rondin, L., Tetienne, J.-P., Hingant, T., Roch, J.-F., Maletinsky, P., Jacques, V.: Magnetometry
with nitrogen-vacancy defects in diamond. Rep. Progress Phys. 77, 056503 (2014)
68. Casola, F., van der Sar, T., Yacoby, A.: Probing condensed matter physics with magnetometry
based on nitrogen-vacancy centres in diamond. Nat. Rev. Mater. 3, 17088 (2018)
69. Chernobrod, B.M., Berman, G.P.: Spin microscope based on optically detected magnetic
resonance. J. Appl. Phys. 97(1), 014903 (2005)
70. Zhou, T.X., Stöhr, R.J., Yacoby, A.: Scanning diamond NV center probes compatible with
conventional AFM technology. Appl. Phys. Lett. 111(16), 163106 (2017)
71. Rondin, L., Tetienne, J.-P., Rohart, S., Thiaville, A., Hingant, T., Spinicelli, P., Roch, J.-
F., Jacques, V.: Stray-field imaging of magnetic vortices with a single diamond spin. Nat.
Commun. 4, 2279 (2013)
72. Grinolds, M.S., Hong, S., Maletinsky, P., Luan, L., Lukin, M.D., Walsworth, R.L., Yacoby, A.:
Nanoscale magnetic imaging of a single electron spin under ambient conditions. Nat. Phys.
9, 215–219 (2013)
73. Tetienne, J.-P., Hingant, T., Kim, J.-V., Diez, L.H., Adam, J.-P., Garcia, K., Roch, J.-F., Rohart,
S., Thiaville, A., Ravelosona, D., Jacques, V.: Nanoscale imaging and control of domain-
wall hopping with a nitrogen-vacancy center microscope. Science 344(6190), 1366–1369
(2014)
74. Dovzhenko, Y., Casola, F., Schlotter, S., Zhou, T.X., Büttner, F., Walsworth, R.L., Beach,
G.S.D., Yacoby, A.: Magnetostatic twists in room-temperature skyrmions explored by
nitrogen-vacancy center spin texture reconstruction. Nat. Commun. 9, 2712 (2018)
75. Thiel, L., Wang, Z., Tschudin, M.A., Rohner, D., Gutiérrez-Lezama, I., Ubrig, N., Gibertini,
M., Giannini, E., Morpurgo, A.F., Maletinsky, P.: Probing magnetism in 2D materials at the
nanoscale with single-spin microscopy. Science 364(6444), 973–976 (2019)
76. Zhou, T.X., Carmiggelt, J.J., Gächter, L.M., Esterlis, I., Sels, D., Stöhr, R.J., Du, C.,
Fernandez, D., Rodriguez-Nieva, J.F., Büttner, F., Demler, E., Yacoby, A.: A magnon
scattering platform (2020)
77. Gross, I., Akhtar, W., Garcia, V., Martínez, L.J., Chouaieb, S., Garcia, K., Carrétéro, C.,
Barthélémy, A., Appel, P., Maletinsky, P., Kim, J.-V., Chauleau, J.Y., Jaouen, N., Viret, M.,
Bibes, M., Fusil, S., Jacques, V.: Real-space imaging of non-collinear antiferromagnetic order
with a single-spin magnetometer. Nature 549, 252–256 (2017)
78. Bode, M.: Spin-polarized scanning tunnelling microscopy. Rep. Progress Phys. 66, 523–582
(2003)
79. Wiesendanger, R.: Spin mapping at the nanoscale and atomic scale. Rev. Mod. Phys. 81,
1495–1550 (2009)
80. Suzuki, Y., Nabhan, W., Tanaka, K.: Magnetic domains of cobalt ultrathin films observed with
a scanning tunneling microscope using optically pumped GaAs tips. Appl. Phys. Lett. 71(21),
3153–3155 (1997)
1250 R. M. Reeve et al.

81. Terada, Y., Yoshida, S., Takeuchi, O., Shigekawa, H.: Real-space imaging of transient carrier
dynamics by nanoscale pump–probe microscopy. Nat. Photon. 4, 869–874 (2010)
82. Alvarado, S.F., Renaud, P.: Observation of spin-polarized-electron tunneling from a ferro-
magnet into GaAs. Phys. Rev. Lett. 68, 1387–1390 (1992)
83. Mühlenberend, S., Gruyters, M., Berndt, R.: Plasmon-mediated circularly polarized lumines-
cence of GaAs in a scanning tunneling microscope. Appl. Phys. Lett. 107(24), 241110 (2015)
84. Schlickum, U., Janke-Gilman, N., Wulfhekel, W., Kirschner, J.: Step-induced frustration of
antiferromagnetic order in Mn on Fe(001). Phys. Rev. Lett. 92, 107203 (2004)
85. Prokop, J., Kukunin, A., Elmers, H.J.: Magnetic anisotropies and coupling mechanisms in
Fe/Mo(110) nanostripes. Phys. Rev. Lett. 95, 187202 (2005)
86. Oka, H., Ignatiev, P.A., Wedekind, S., Rodary, G., Niebergall, L., Stepanyuk, V.S., Sander, D.,
Kirschner, J.: Spin-dependent quantum interference within a single magnetic nanostructure.
Science 327(5967), 843–846 (2010)
87. Pratzer, M., Elmers, H.J., Bode, M., Pietzsch, O., Kubetzka, A., Wiesendanger, R.: Atomic-
scale magnetic domain walls in quasi-one-dimensional Fe nanostripes. Phys. Rev. Lett. 87,
127201 (2001)
88. Prokop, J., Kukunin, A., Elmers, H.J.: Spin-polarized scanning tunneling microscopy and
spectroscopy of ultrathin Fe/Mo(110) films using W/Au/Co tips. Phys. Rev. B 73, 014428
(2006)
89. Wortmann, D., Heinze, S., Kurz, P., Bihlmayer, G., Blügel, S.: Resolving complex atomic-
scale spin structures by spin-polarized scanning tunneling microscopy. Phys. Rev. Lett. 86,
4132–4135 (2001)
90. Bardeen, J.: Tunnelling from a many-particle point of view. Phys. Rev. Lett. 6, 57–59 (1961)
91. Schwöbel, J., Fu, Y., Brede, J., Dilullo, A., Hoffmann, G., Klyatskaya, S., Ruben, M.,
Wiesendanger, R.: Real-space observation of spin-split molecular orbitals of adsorbed single-
molecule magnets. Nat. Commun. 3, 953 (2012)
92. Methfessel, T., Steil, S., Baadji, N., Großmann, N., Koffler, K., Sanvito, S., Aeschlimann,
M., Cinchetti, M., Elmers, H.J.: Spin scattering and spin-polarized hybrid interface states at a
metal-organic interface. Phys. Rev. B 84, 224403 (2011)
93. Heinze, S., von Bergmann, K., Menzel, M., Brede, J., Kubetzka, A., Wiesendanger, R.,
Bihlmayer, G., Blügel, S.: Spontaneous atomic-scale magnetic skyrmion lattice in two
dimensions. Nat. Phys. 7, 713–718 (2011)
94. Romming, N., Kubetzka, A., Hanneken, C., von Bergmann, K., Wiesendanger, R.: Field-
dependent size and shape of single magnetic skyrmions. Phys. Rev. Lett. 114, 177203 (2015)
95. Leonov, A.O., Monchesky, T.L., Romming, N., Kubetzka, A., Bogdanov, A.N., Wiesendan-
ger, R.: The properties of isolated chiral skyrmions in thin magnetic films. New J. Phys. 18,
065003 (2016)
96. Romming, N., Hanneken, C., Menzel, M., Bickel, J.E., Wolter, B., von Bergmann, K.,
Kubetzka, A., Wiesendanger, R.: Writing and deleting single magnetic skyrmions. Science
341(6146), 636–639 (2013)
97. Martin, Y., Wickramasinghe, H.K.: Magnetic imaging by “force microscopy” with 1000 Å
resolution. Appl. Phys. Lett. 50(20), 1455–1457 (1987)
98. Schwarz, A., Wiesendanger, R.: Magnetic sensitive force microscopy. Nano Today 3(1), 28–
39 (2008)
99. Hartmann, U.: Magnetic force microscopy. Ann. Rev. Mater. Sci. 29(1), 53–87 (1999)
100. Rugar, D., Mamin, H.J., Guethner, P., Lambert, S.E., Stern, J.E., McFadyen, I., Yogi, T.:
Magnetic force microscopy: general principles and application to longitudinal recording
media. J. Appl. Phys. 68(3), 1169–1183 (1990)
101. Hug, H.J., Stiefel, B., van Schendel, P.J.A., Moser, A., Hofer, R., Martin, S., Güntherodt,
H.-J., Porthun, S., Abelmann, L., Lodder, J.C., Bochi, G., O’Handley, R.C.: Quantitative
magnetic force microscopy on perpendicularly magnetized samples. J. Appl. Phys. 83(11),
5609–5620 (1998)
24 Magnetic Imaging and Microscopy 1251

102. Belova, L.M., Hellwig, O., Dobisz, E., Dan Dahlberg, E.: Rapid preparation of electron beam
induced deposition Co magnetic force microscopy tips with 10 nm spatial resolution. Rev.
Sci. Instrum. 83(9), 093711 (2012)
103. Baćani, M., Marioni, M.A., Schwenk, J., Hug, H.J.: How to measure the local dzyaloshinskii-
moriya interaction in skyrmion thin-film multilayers. Sci. Rep. 9, 3114 (2019)
104. Stoll, H., Noske, M., Weigand, M., Richter, K., Krüger, B., Reeve, R.M., Hänze, M., Adolff,
C.F., Stein, F.-U., Meier, G., Kläui, M., Schütz, G.: Imaging spin dynamics on the nanoscale
using x-ray microscopy. Front. Phys. 3, 26 (2015)
105. Fischer, P., Ohldag, H.: X-rays and magnetism. Rep. Progress Phys. 78, 094501 (2015)
106. Boulle, O., Vogel, J., Yang, H., Pizzini, S., de Souza Chaves, D., Locatelli, A., Menteş, T.O.,
Sala, A., Buda-Prejbeanu, L.D., Klein, O., Belmeguenai, M., Roussigné, Y., Stashkevich,
A., Chérif, S.M., Aballe, L., Foerster, M., Chshiev, M., Auffret, S., Miron, I.M., Gaudin,
G.: Room-temperature chiral magnetic skyrmions in ultrathin magnetic nanostructures. Nat.
Nanotechnol. 12, 830–830 (2017)
107. Zeissler, K., Finizio, S., Shahbazi, K., Massey, J., Ma’Mari, F.A., Bracher, D.M., Kleibert, A.,
Rosamond, M.C., Linfield, E.H., Moore, T.A., Raabe, J., Burnell, G., Marrows, C.H.: Discrete
Hall resistivity contribution from Néel skyrmions in multilayer nanodiscs. Nat. Nanotechnol.
13, 1161–1166 (2018)
108. Woo, S., Song, K.M., Han, H.-S., Jung, M.-S., Im, M.-Y., Lee, K.-S., Song, K.S., Fischer, P.,
Hong, J.-I., Choi, J.W., Min, B.-C., Koo, H.C., Chang, J.: Spin-orbit torque-driven skyrmion
dynamics revealed by time-resolved x-ray microscopy. Nat. Commun. 8, 15573 (2017)
109. Büttner, F., Lemesh, I., Schneider, M., Pfau, B., Günther, C.M., Hessing, P., Geilhufe,
J., Caretta, L., Engel, D., Krüger, B., Viefhaus, J., Eisebitt, S., Beach, G.S.D.: Field-
free deterministic ultrafast creation of magnetic skyrmions by spin–orbit torques. Nat.
Nanotechnol. 12, 1040–1044 (2017)
110. Schütz, G., Wagner, W., Wilhelm, W., Kienle, P., Zeller, R., Frahm, R., Materlik, G.:
Absorption of circularly polarized x rays in iron. Phys. Rev. Lett. 58, 737–740 (1987)
111. Stöhr, J., Siegmann, H.C.: Magnetism from Fundamentals to Nanoscale Dynamics. Springer,
Berlin/Heidelberg (2006)
112. Fan, T., Grychtol, P., Knut, R., Hernández-García, C., Hickstein, D.D., Zusin, D., Gentry, C.,
Dollar, F.J., Mancuso, C.A., Hogle, C.W., Kfir, O., Legut, D., Carva, K., Ellis, J.L., Dorney,
K.M., Chen, C., Shpyrko, O.G., Fullerton, E.E., Cohen, O., Oppeneer, P.M., Milošević, D.B.,
Becker, A., Jaroń-Becker, A.A., Popmintchev, T., Murnane, M.M., Kapteyn, H.C.: Bright
circularly polarized soft x-ray high harmonics for x-ray magnetic circular dichroism. Proc.
Natl. Acad. Sci. 112(46), 14206–14211 (2015)
113. Willems, F., Smeenk, C.T.L., Zhavoronkov, N., Kornilov, O., Radu, I., Schmidbauer, M.,
Hanke, M., von Korff Schmising, C., Vrakking, M.J.J., Eisebitt, S.: Probing ultrafast spin
dynamics with high-harmonic magnetic circular dichroism spectroscopy. Phys. Rev. B 92,
220405 (2015)
114. Kfir, O., Zayko, S., Nolte, C., Sivis, M., Möller, M., Hebler, B., Arekapudi, S.S.P.K., Steil, D.,
Schäfer, S., Albrecht, M., Cohen, O., Mathias, S., Ropers, C.: Nanoscale magnetic imaging
using circularly polarized high-harmonic radiation. Sci. Adv. 3(12), eaao4641 (2017)
115. Fischer, P., Im, M.-Y., Baldasseroni, C., Bordel, C., Hellman, F., Lee, J.-S., Fadley, C.S.:
Magnetic imaging with full-field soft x-ray microscopies. J. Electron Spectrosc. Relat.
Phenom. 189, 196–205 (2013)
116. Sakdinawat, A., Attwood, D.: Nanoscale x-ray imaging. Nat. Photon. 4, 840–848 (2010)
117. Kilcoyne, A.L.D., Tyliszczak, T., Steele, W.F., Fakra, S., Hitchcock, P., Franck, K., Anderson,
E., Harteneck, B., Rightor, E.G., Mitchell, G.E., Hitchcock, A.P., Yang, L., Warwick, T., Ade,
H.: Interferometer-controlled scanning transmission x-ray microscopes at the advanced light
source. J. Synchrotron Radiat. 10(2), 125–136 (2003)
118. Shi, X., Fischer, P., Neu, V., Elefant, D., Lee, J.C.T., Shapiro, D.A., Farmand, M., Tyliszczak,
T., Shiu, H.-W., Marchesini, S., Roy, S., Kevan, S.D.: Soft x-ray ptychography studies of
1252 R. M. Reeve et al.

nanoscale magnetic and structural correlations in thin SmCo5 films. Appl. Phys. Lett. 108(9),
094103 (2016)
119. Donnelly, C., Finizio, S., Gliga, S., Holler, M., Hrabec, A., Odstrčil, M., Mayr, S., Scagnoli,
V., Heyderman, L.J., Guizar-Sicairos, M., Raabe, J.: Time-resolved imaging of three-
dimensional nanoscale magnetization dynamics. Nat. Nanotechnol. 15, 356–360 (2020)
120. Li, W., Bykova, I., Zhang, S., Yu, G., Tomasello, R., Carpentieri, M., Liu, Y., Guang, Y.,
Gräfe, J., Weigand, M., Burn, D.M., van der Laan, G., Hesjedal, T., Yan, Z., Feng, J., Wan,
C., Wei, J., Wang, X., Zhang, X., Xu, H., Guo, C., Wei, H., Finocchio, G., Han, X., Schütz,
G.: Anatomy of skyrmionic textures in magnetic multilayers. Adv. Mater. 31(14), 1807683
(2019)
121. Stöhr, J., Wu, Y., Hermsmeier, B.D., Samant, M.G., Harp, G.R., Koranda, S., Dunham, D.,
Tonner, B.P.: Element-specific magnetic microscopy with circularly polarized x-rays. Science
259(5095), 658–661 (1993)
122. Cheng, X.M., Keavney, D.J.: Studies of nanomagnetism using synchrotron-based x-ray
photoemission electron microscopy (X-PEEM). Rep. Progress Phys. 75, 026501 (2012)
123. Rhensius, J., Heyne, L., Backes, D., Krzyk, S., Heyderman, L.J., Joly, L., Nolting, F.,
Kläui, M.: Imaging of domain wall inertia in permalloy half-ring nanowires by time-resolved
photoemission electron microscopy. Phys. Rev. Lett. 104, 067201 (2010)
124. Chapman, H.N., Nugent, K.A.: Coherent lensless x-ray imaging. Nat. Photon. 4, 833–839
(2010)
125. Donnelly, C., Guizar-Sicairos, M., Scagnoli, V., Gliga, S., Holler, M., Raabe, J., Heyderman,
L.J.: Three-dimensional magnetization structures revealed with x-ray vector nanotomography.
Nature 547(7663), 328–331 (2017)
126. Thibault, P., Dierolf, M., Menzel, A., Bunk, O., David, C., Pfeiffer, F.: High-resolution
scanning x-ray diffraction microscopy. Science 321(5887), 379–382 (2008)
127. Thibault, P., Menzel, A.: Reconstructing state mixtures from diffraction measurements.
Nature 494, 68–71 (2013)
128. Eisebitt, S., Lüning, J., Schlotter, W.F., Lörgen, M., Hellwig, O., Eberhardt, W., Stöhr, J.:
Lensless imaging of magnetic nanostructures by x-ray spectro-holography. Nature 432, 885–
888 (2004)
129. Flewett, S., Günther, C.M., von Korff Schmising, C., Pfau, B., Mohanty, J., Büttner, F.,
Riemeier, M., Hantschmann, M., Kläui, M., Eisebitt, S.: Holographically aided iterative phase
retrieval. Opt. Express 20, 29210–29216 (2012)
130. Turner, J.J., Huang, X., Krupin, O., Seu, K.A., Parks, D., Kevan, S., Lima, E., Kisslinger, K.,
McNulty, I., Gambino, R., Mangin, S., Roy, S., Fischer, P.: X-ray diffraction microscopy of
magnetic structures. Phys. Rev. Lett. 107, 033904 (2011)
131. Scherz, A., Schlotter, W.F., Chen, K., Rick, R., Stöhr, J., Lüning, J., McNulty, I., Günther, C.,
Radu, F., Eberhardt, W., Hellwig, O., Eisebitt, S.: Phase imaging of magnetic nanostructures
using resonant soft x-ray holography. Phys. Rev. B 76, 214410 (2007)
132. Chapman, H.N., Barty, A., Bogan, M.J., Boutet, S., Frank, M., Hau-Riege, S.P., Marchesini,
S., Woods, B.W., Bajt, S., Benner, W.H., London, R.A., Plönjes, E., Kuhlmann, M., Treusch,
R., Düsterer, S., Tschentscher, T., Schneider, J.R., Spiller, E., Möller, T., Bostedt, C., Hoener,
M., Shapiro, D.A., Hodgson, K.O., van der Spoel, D., Burmeister, F., Bergh, M., Caleman,
C., Huldt, G., Seibert, M.M., Maia, F.R.N.C., Lee, R.W., Szöke, A., Timneanu, N., Hajdu, J.:
Femtosecond diffractive imaging with a soft-x-ray free-electron laser. Nat. Phys. 2, 839–843
(2006)
133. Flewett, S., Schaffert, S., Mohanty, J., Guehrs, E., Geilhufe, J., Günther, C.M., Pfau, B.,
Eisebitt, S.: Method for single-shot coherent diffractive imaging of magnetic domains. Phys.
Rev. Lett. 108, 223902 (2012)
134. Büttner, F., Moutafis, C., Schneider, M., Krüger, B., Günther, C.M., Geilhufe, J., Schmising,
C.V.K., Mohanty, J., Pfau, B., Schaffert, S., Bisig, A., Foerster, M., Schulz, T., Vaz, C.A.F.,
Franken, J.H., Swagten, H.J.M., Kläui, M., Eisebitt, S.: Dynamics and inertia of skyrmionic
spin structures. Nat. Phys. 11, 225–228 (2015)
24 Magnetic Imaging and Microscopy 1253

135. Stefan, H.: Photoelectron Spectroscopy: Principles and Applications. Springer, Berlin/Lon-
don (2011)
136. Schönhense, G., Medjanik, K., Elmers, H.-J.: Space-, time- and spin-resolved photoemission.
J. Electron Spectrosc. Relat. Phenom. 200, 94–118 (2015)
137. Fujiwara, H., Kiss, T., Wakabayashi, Y.K., Nishitani, Y., Mori, T., Nakata, Y., Kitayama, S.,
Fukushima, K., Ikeda, S., Fuchimoto, H., et al.: Soft x-ray angle-resolved photoemission with
micro-positioning techniques for metallic v2o3. J. Synchrotron Radiat. 22(3), 776–780 (2015)
138. Tusche, C., Krasyuk, A., Kirschner, J.: Spin resolved bandstructure imaging with a high
resolution momentum microscope. Ultramicroscopy 159, 520–529 (2015)
139. Hendee, W.R.: Physics and applications of medical imaging. Rev. Mod. Phys. 71, S444–S450
(1999)
140. Kathiravan, S., Kanakaraj, J.: A review on potential issues and challenges in MR imaging.
Sci. World J. 2013, 783715 (2013)
141. Odaibo, S.G.: Quantum Mechanics and the MRI Machine. Symmetry Seed Books, Arlington
(2012)
142. Plewes, D.B., Kucharczyk, W.: Physics of MRI: a primer. J. Magn. Reson. Imaging 35, 1038–
1054 (2012)
143. Edelman, R.R., Warach, S.: Magnetic resonance imaging. N. Engl. J. Med. 328(10), 708–716
(1993)
144. Blundell, S.J.: Magnetism in Condensed Matter. Oxford University Press, Oxford (2014)
145. Jung, B.A., Weigel, M.: Spin echo magnetic resonance imaging. J. Magn. Reson. Imaging
37(4), 805–817 (2013)
146. Suits, B.A.: Nuclear quadrupole resonance spectroscopy. In: Vij, D.R. (ed.) Handbook of
Applied Solid State Spectroscopy. Springer, Berlin/Heidelberg/New York (2006)
147. Robert, H., Pusiol, D.: Two-dimensional rotating-frame NQR imaging. J. Magn. Reson.
127(1), 109–114 (1997)
148. Nickel, P., Robert, H., Kimmich, R., Pusiol, D.: NQR method for stress and pressure imaging.
J. Magn. Reson. Ser. A 111(2), 191–194 (1994)
149. Hari, R., Salmelin, R.: Magnetoencephalography: from squids to neuroscience. NeuroImage
61(2), 386–396 (2012)
150. Braeutigam, S.: Magnetoencephalography: fundamentals and established and emerging
clinical applications in radiology. ISRN Radiology 2013, 1–18 (2013)
151. Hämäläinen, M., Hari, R., Ilmoniemi, R.J., Knuutila, J., Lounasmaa, O.V.:
Magnetoencephalography—theory, instrumentation, and applications to noninvasive
studies of the working human brain. Rev. Mod. Phys. 65, 413–497 (1993)
152. Saunus, C., Raphael Bindel, J., Pratzer, M., Morgenstern, M.: Versatile scanning tunneling
microscopy with 120 ps time resolution. Appl. Phys. Lett. 102(5), 051601 (2013)

Robert M. Reeve received his PhD from the University of Cam-


bridge for his work on surface magnetism and spin-dependent
electron scattering from ultrathin magnetic films. Currently, he
researches at Johannes Gutenberg University Mainz with projects
focusing on magnetism, spin transport, and spin dynamics. His
work combines transport and magnetic imaging to study the
dynamics of spin states in mesoscopic systems.
1254 R. M. Reeve et al.

Hans-Joachim Elmers received his PhD in physics from the


University of Clausthal, Germany, in 1989. He habilitated in
experimental physics at the University of Clausthal, Germany, in
1995. Since 1998, he is a professor in experimental physics at the
University of Mainz, Germany. His research focuses on ultrathin
magnetic films and spin effects at interfaces.

Felix Büttner received his PhD from the University of Mainz.


After a postdoc at MIT, he joined Helmholtz-Zentrum Berlin with
an independent Young Investigator Grant. His work focuses on
x-ray imaging, in particular x-ray holography of the dynamics
of topological magnetic spin textures with ultrahigh spatial and
temporal resolution, as well the design of materials and devices
to this end.

Mathias Kläui received his PhD from Cambridge University.


He worked at IBM Zurich and the University of Konstanz before
being appointed Associate Professor jointly at EPFL and PSI. In
2011, he became full Professor at Johannes Gutenberg University
Mainz, and since 2017, he is Adjunct Professor at NTNU in
Norway. He works on magnetism, spin dynamics, and low-
dimensional systems.
Magnetic Scattering
25
Jeffrey W. Lynn and Bernhard Keimer

Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1256
Magnetic Neutron Diffraction Technique . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1257
Polarized Neutron Techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1260
Polarized Neutron Reflectometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1261
Resonant Magnetic X-ray Diffraction Technique . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1262
Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1265
Inelastic Neutron Scattering Technique . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1265
Resonant Inelastic X-ray Scattering Technique . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1267
Magnetic Diffraction Examples with Neutrons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1268
Magnetic Diffraction Examples with X-rays . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1276
Spin Dynamics with Neutrons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1282
Spin Dynamics with RIXS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1287
Facilities and Online Information . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1289
Summary and Future Directions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1291
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1292

Abstract

The present chapter reviews current neutron and X-ray scattering techniques
employed to elucidate the magnetic structures and spin dynamics of mag-
netic materials. Both techniques provide measurements as a function of the
energy and the momentum transferred from the spin system to the probe
particles, in terms of five-dimensional data sets as a function of various ther-

J. W. Lynn ()
NIST Center for Neutron Research, National Institute of Standards and Technology,
Gaithersburg, MD, USA
e-mail: jeffrey.lynn@nist.gov
B. Keimer
Max-Planck Institute for Solid State Research, Stuttgart, Germany
e-mail: b.keimer@fkf.mpg.de

© Springer Nature Switzerland AG 2021 1255


J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_26
1256 J. W. Lynn and B. Keimer

modynamic fields at the control of the experimenter. These scattering tech-


niques yield fundamental information about the equal-time correlations such
the magnetic configuration and symmetry, as well as the dynamics that evince
the exchange interactions for prototypical systems that behave as linear, pla-
nar, or three-dimensional systems. Historically, neutron scattering has been
the magnetic scattering technique of choice for such investigations, but the
extraordinary advances in resonant X-ray scattering techniques have enabled
new types of magnetic scattering measurements. The type of information
obtained with the two techniques is largely complementary and depends on
the interests of the investigators. We discuss these possibilities and provide
numerous examples of the techniques applied to different classes of magnetic
systems.

Introduction

Scattering measurements provide essential information about the intrinsic electronic


interactions in magnetic materials. Traditionally neutron scattering has been the
unique magnetic scattering tool, [1, 2] but that situation has changed recently
following remarkable advances in resonant X-ray scattering techniques [3]. Both
techniques collect data as functions of the energy and the momentum transferred
from the spin system to the neutron or photon beam. The resulting five-dimensional
data sets serve as powerful probes of magnetic materials. Elastic scattering eluci-
dates the magnetic configuration, direction of the spins, symmetry of the magnetic
state, spatial distribution of the magnetization density, and dependence of the
order parameter on thermodynamic fields such as temperature, pressure, magnetic
and electric fields. Inelastic scattering determines the energies of the fundamental
excitations, which can be used to elucidate the nature and strength of the exchange
interactions.
The information provided by X-ray and neutron scattering is largely com-
plementary [4]. For instance, resonant elastic X-ray scattering can be used to
measure magnetic order parameters in an element-specific manner by tuning the
X-ray energy to an electronic transition, and the enormous photon flux at modern
synchrotron facilities allows measurements of very small samples and films as thin
as a single unit cell [5, 6]. The magnetic dynamics of such samples can be explored
with Resonant Inelastic X-ray Scattering (RIXS), [7] and pump-probe techniques
open new avenues of research into nonequilibrium phenomena. Elastic neutron
scattering, on the other hand, provides quantitative information about the magnitude
of magnetic order parameters that is difficult to obtain with X-rays, and the energy
resolution offered by inelastic magnetic neutron scattering is orders-of-magnitude
finer than can be currently achieved in RIXS experiments. Both techniques can also
measure the lattice structure and dynamics, as reviewed elsewhere in this volume,
and those cross sections can be uniquely distinguished from magnetic scattering by
polarization techniques [8, 9].
25 Magnetic Scattering 1257

In this chapter, we describe how to employ neutron and X-ray scattering to


explore the magnetism of materials, paying particular attention to the complemen-
tarity of both techniques.

Magnetic Neutron Diffraction Technique

Magnetic neutron scattering originates from the neutron’s magnetic dipole moment.
As a spin-½ fermion, the neutron carries a magnetic dipole moment of −1.913
nuclear magnetons that interacts with the unpaired electrons in the sample, either
through the dipole moment associated with an electron’s spin or via the orbital
motion of the electron. The strength of this magnetic interaction is comparable
to the neutron–nuclear interaction. The magnetic scattering cross section reveals
the magnetic structure and dynamics of materials over wide ranges of length scale
and energy. Magnetic neutron scattering plays a central role in determining and
understanding the microscopic properties of a vast variety of magnetic systems –
from the fundamental nature, symmetry, and dynamics of magnetically ordered
materials to elucidating the magnetic characteristics essential in technological
applications.
One traditional role of magnetic neutron scattering has been the measurement of
magnetic Bragg intensities in the magnetically ordered regime. Such measurements
can be used to determine the spatial arrangement and directions of the atomic
magnetic moments, the atomic magnetization density of the individual atoms in
the material, and the value of the ordered moments as a function of external
parameters such as temperature, pressure, and applied magnetic or electric fields.
These types of measurements can be carried out on single crystals, powders, thin
films, and artificially grown multilayers, and often the information collected can be
obtained by no other experimental technique. For magnetic phenomena that occur
over length scales that are large compared to atomic distances, the technique of
magnetic Small Angle Neutron Scattering (SANS) can be applied. This is an ideal
technique to explore domain structures, long wavelength oscillatory magnetic states,
vortex structures in superconductors, skyrmions, nanomagnets, and other spatial
variations of the magnetization density on length scales from 1 to 1000 nm. Another
specialized technique is neutron reflectometry, which can be used to investigate the
magnetization profile in the near-surface regime of single crystals, as well as the
magnetization density of thin films and multilayers. This particular technique has
enjoyed dramatic growth during the last decade or so due to the rapid advancement
of atomic-layer deposition capabilities.
The cross section for magnetic Bragg scattering can be written as Ref. [10]
 2
γ e2
IM (ghkl ) = C Mg A (θB ) |FM (ghkl )|2 (1)
2mc2

where IM is the integrated intensity for the magnetic Bragg reflection located at the
reciprocal lattice vector ghkl , the neutron-electron coupling constant in parentheses
1258 J. W. Lynn and B. Keimer

is −0.27 × 10−12 cm, C is an instrumental constant which includes the resolution


of the measurement, A(θ B ) is an angular factor which depends on the method of
measurement (sample angular rotation, θ :2θ scan, etc.), and Mg is the multiplicity
of the reflection (for a powder sample). The magnetic structure factor FM (g) is given
in the general case by Refs. [1, 11]


N
 
FM (ghkl ) = eig·rj ĝ × M j (g) × ĝ e−Wj (2)
j =1

where ĝ is a unit vector in the direction of the reciprocal lattice vector ghkl , Mj (ghkl )
is the vector form factor of the j-th ion located at rj in the unit cell, Wj is the Debye-
Waller factor that accounts for the thermal vibrations of the j-th ion, and the sum is
over all (magnetic) atoms in the unit cell. The triple cross product originates from
the vector nature of the dipole–dipole interaction of the neutron with the electron.
A quantitative calculation of Mj (g) in the general case involves evaluating matrix
elements of both spin and orbital operators [11, 12]. This can be quite a complicated
angular-momentum computation involving all the electron orbitals in the unit cell,
but has the simple result that only the components of the magnetic moment that
are perpendicular to ghkl (or more generally the wave vector K) contribute to the
scattering. Often the atomic spin density is collinear, by which we mean that at
each point in the spatial extent of the electron’s probability distribution, the atomic
magnetization density points in the same direction. In this case the direction of
Mj (g) does not depend on g, and the form factor is just a scalar function, f (g),
which is simply related to the Fourier transform of the magnetization density. The
free-ion form factors have been tabulated for essentially all the magnetic elements
(see, for example, https://www.ill.eu/sites/ccsl/ffacts/ffachtml.html). Note that for
X-ray scattering the form factor for charge scattering corresponds to the Fourier
transform of the total charge density of all the electrons, while in the magnetic
neutron case it is the transform of the “magnetic” electrons only, which are the
electrons whose spins are unpaired. Recalling that a Fourier transform inverts the
relative size of objects, the magnetic form factor typically decreases much more
rapidly with |ghkl | than for the case of X-ray charge scattering since the unpaired
electrons are usually the outermost ones of the ion. This dependence of the scattering
intensity on f (g) is a convenient way to distinguish magnetic cross sections from
nuclear cross sections, where the equivalent of the form factor is just a constant
since the nucleus (≈10−5 Å) looks like a point particle to a thermal/cold neutron
(see the nuclear coherent scattering amplitude b in Eq. (6)).
If in addition to the magnetization density being collinear, the magnetic moments
in the ordered state point along a unique direction (i.e., the magnetic structure is a
ferromagnet, or a simple + − + − type antiferromagnet), then the square of the
magnetic structure factor simplifies to
2


 


2 −Wj ig·rj
|FM (g)| = 1 − ĝ · η̂
2 z
ηj μj fj (g) e e , (3)
j
25 Magnetic Scattering 1259

where η̂ denotes the (common)



direction of the ordered moments and ηj the sign
z
of the moment (±1), μj is the average value of the ordered moment in ther-
  2

modynamic equilibrium at (T, H, P, . . . ), and the orientation factor 1 − ĝ · η̂


represents an average over all possible domains. If the magnetic moments are the
same type, then this expression further simplifies to
2


 
|FM (g)| = 1 − ĝ · η̂
2 2 z 2 2
μ f (g) ηj e ig·r j −Wj
e (4)
.
j

We see from these expressions that neutrons can be used to determine several
important quantities: the location of magnetic atoms in the unit cell and the spatial
distribution of their magnetic electrons; the dependence of <μz > on temperature,
field, pressure, or other thermodynamic variables, which is directly related to the
order parameter for the phase transition (e.g., the sublattice magnetization). Often
the preferred magnetic axis η̂ can also be determined from the relative intensities.
Finally, the scattering can be put on an absolute scale by internal comparison with
the nuclear Bragg intensities IN from the same sample, given by

IN (g) = CMg A (θB ) |FN (g)|2 (5)

with
2


2
|FN (g)| = bj e ig·r j −Wj
e (6)
.
j

Here, bj is the coherent nuclear scattering amplitude for the j-th atom in the unit
cell, and the sum is over all atoms in the unit cell. Typically, the nuclear structure
is known accurately and FN can be calculated, whereby the saturated value of the
magnetic moment in Bohr magnetons can be obtained.
There are several ways that magnetic Bragg scattering can be distinguished from
the nuclear scattering from the structure. Above the magnetic ordering temperature
all Bragg peaks are nuclear (structural) in origin, while as the temperature drops
below the ordering temperature the intensities of the magnetic Bragg peaks rapidly
develop, and for unpolarized neutrons the nuclear and magnetic intensities simply
add. If these new Bragg peaks occur at positions that are distinct from the nuclear
reflections, then it is straightforward to distinguish magnetic from nuclear scattering.
In the case of a ferromagnet, however, or for some antiferromagnets which contain
two or more magnetic atoms in the chemical unit cell, these Bragg peaks can occur
at the same position. One standard technique for identifying the magnetic Bragg
scattering is to make one diffraction measurement in the paramagnetic state well
above the ordering temperature, and another in the ordered state at the lowest
temperature possible, and then subtract the two sets of data. In the paramagnetic
1260 J. W. Lynn and B. Keimer

state the (free ion) diffuse magnetic scattering is given by Refs. [1, 10]

 2
2 γ e2
IP ara = C 2
peff f (K)2 (7)
3 2mc2

where peff is the effective magnetic moment (= g[J(J + 1)]1/2 for a free ion). This
is a magnetic incoherent cross section, and the only angular dependence is through
the magnetic form factor f (K). Hence this scattering looks like “background.”
There is a sum rule on the magnetic scattering in the system, though, and in the
ordered state this diffuse scattering shifts into the coherent magnetic Bragg peaks
and magnetic excitations. A subtraction of the high temperature data in (7) from
the data obtained at low temperature (1) will then yield the magnetic Bragg peaks,
on top of a deficit (negative) of scattering away from the Bragg peaks due to the
disappearance of the diffuse paramagnetic scattering in the ordered state. On the
other hand, all the nuclear cross sections usually do not change significantly with
temperature (apart from the Debye-Waller factor e−2W ), and hence drop out in the
subtraction. A related subtraction technique is to apply a large magnetic field in
the paramagnetic state, to induce a net (ferromagnetic-like) moment. The zero field
(nuclear) diffraction pattern can then be subtracted from the high-field pattern to
obtain the induced-moment diffraction pattern.

Polarized Neutron Techniques

When the neutron beam impinges on a sample in a well-defined polarization state,


then the nuclear and magnetic scattering cross sections that originate from the
sample interfere coherently, in contrast to being separate cross sections like (1) and
(5) where magnetic and nuclear intensities just add. Polarized neutron diffraction
measurements with polarization analysis of the scattered neutrons can be used
to establish unambiguously which peaks are magnetic, which are nuclear, and
more generally to separate the magnetic and nuclear scattering at Bragg positions
where there are both nuclear and magnetic contributions. The standard polarization
analysis technique is straightforward in principle [9, 13]. Nuclear coherent Bragg
scattering never causes a reversal, or spin-flip, of the neutron spin direction upon
scattering. Thus the nuclear peaks will only be observed in the non-spin-flip
scattering geometry. We denote this configuration as (+ +), where the incident
spin of the neutron is “up” spin and remains in the up state after scattering. Non-
spin-flip scattering also occurs if the incident neutron is in the “down” state, and
remains in the down state after scattering (denoted (− −)). The magnetic cross
sections, on the other hand, depend on the relative orientation of the neutron
polarization P and the reciprocal lattice vector g. In the configuration where P⊥g,
typically half the magnetic Bragg scattering involves a reversal of the neutron
spin (denoted by (− +) or (+ −)), and half does not; the details depend on the
specific Hamiltonian describing the magnetism. Thus for an isotropic Heisenberg-
25 Magnetic Scattering 1261

type model the magnetic contribution to the reflection consists of the spin-flip (− +)
and non-spin-flip (+ +) intensities of equal intensity. For the case where P||g, all the
magnetic scattering is spin-flip. Hence for a pure magnetic Bragg reflection where
(Sx ,Sy ,Sz ) are active, the spin-flip scattering should be twice as strong as for the P⊥g
configuration.
The arrangement of having P||g or P⊥g provides an experimental simplification
and hence data that are straightforward to interpret. More generally, however, P and
g can have any relative angle. This more general technique of neutron polarimetry
is more difficult to realize experimentally and can complicate the interpretation of
the data, but can provide additional details about the magnetic structure that cannot
be obtained otherwise [14].

Polarized Neutron Reflectometry

If neutrons are incident on a surface at (very small) grazing angles the scattering
can be cast in the form of a neutron “optical potential,” analogous to photons in
optical fibers. For most materials the wavelength-dependent index of refraction for
neutrons (and x-rays) n is slightly less than unity, so that at sufficiently small angles
of incidence the scattering can be described by the one-dimensional Schrödinger
equation and the neutrons undergo total external reflection – the basis for neutron
guides. For a simple material with a net magnetization, interference between nuclear
and magnetic scattering leads to the following expression for n [10, 15]:

    1/2
± λ2 γ e2
n = 1− N b± μ , (8)
π 2mc2

where N is the number density of the material and < μ > is the average moment.
The magnetic form factor is unity since we are scattering at very small angles. Note
that N b is the nuclear scattering length density for the material, and the magnetic
term is the magnetic scattering length density. The critical angle below which we
have mirror reflection is given by

    1/2  2    1/2
λ2 γ e2 ∼ λ γ e2
θC = arcsin N b± μ = N b ± μ ,
π 2mc2 π 2mc2
(9)

where ± denotes the two polarization states of the neutron. Above the critical angle
the neutrons penetrate the surface, and Fourier transforming the scattering provides
a quantitative measure of the structural profile and magnetic profile of the material.
For thin films and multilayers the layers, substrate, and front and back surfaces
produce interference effects that provide a standard and very powerful technique for
determining the properties of a wide variety of magnetic materials [16, 17].
1262 J. W. Lynn and B. Keimer

Resonant Magnetic X-ray Diffraction Technique

Magnetic X-ray scattering was first demonstrated off resonance, that is, with
photons that were not tuned to any absorption edge of the material under study.
However, the nonresonant magnetic X-ray scattering cross section is so small that
this technique is not useful for magnetic structure determination. Magnetic X-
ray scattering has only risen to prominence when synchrotron radiation enabled
experiments with photons tuned to X-ray absorption edges, where the resonant cross
section can be enhanced by several orders of magnitude [5, 6]. The enhancement is
greatest when the partially occupied valence shell is reached by an electric dipole-
allowed transition, that is, at the L2,3 -absorption edges of transition metals with
valence d-electrons, and at the M4,5 -absorption edges of lanthanides or actinides
with valence f-electrons. Magnetic X-ray scattering is then activated by the strong
core-hole spin-orbit coupling in the intermediate state, prior to reemission of the
photon.
From an instrumental perspective, one can group magnetic X-ray scattering
experiments into three categories, depending on the photon energy E required
to reach the respective absorption edges, namely, soft (E < 1 keV), intermediate
(1 ≤ E ≤ 5 keV), and hard (E > 5 keV). Whereas soft X-ray experiments use
gratings to monochromate the synchrotron radiation, intermediate and hard x-ray
experiments are performed with single-crystal monochromators. Because of air
absorption, soft and intermediate X-ray experiments are carried out under vacuum
conditions. The soft and intermediate X-ray ranges comprise the L-edges of 3d (4d)
metals and the M-edges of 4f- (5f-) electron systems, respectively. Experiments
at the dipole-active L-edges of 5d metals are carried out with hard X-rays, as are
experiments at the K-absorption edges of d-electron systems and L-absorption edges
of f-electron systems where the resonant enhancement of the magnetic cross section
is weaker.
Unlike neutron scattering, resonant magnetic X-ray scattering experiments
require photons with a specific energy, so that only the direction and not the
magnitude of the photon momentum is adjustable. Momentum conservation yields
kinematic constraints that are particularly severe for soft X-ray experiments on
the important class of 3d metal compounds, where simple antiferromagnetic
Bragg reflections characteristic of a doubled crystallographic unit cell cannot be
reached in many cases (Fig. 1 [7]). Magnetic order with larger periodicities (and
correspondingly shorter reciprocal lattice vectors) can be studied by resonant X-ray
diffraction, but dynamical diffraction effects can be important (see the example
below). For resonant X-ray diffraction with intermediate and hard X-rays (Fig. 1),
these constraints do not apply.
In contrast to magnetic neutron scattering which is generally straightforward
to interpret, a complete quantitative calculation of the magnetic X-ray scattering
cross section requires numerical electronic structure calculations that describe the
many-body correlations in the intermediate state. In many cases, however, one is
25 Magnetic Scattering 1263

Fig. 1 (Top)
energy/density-of-states
diagram illustrating RIXS
with photons near the
L-absorption edges of 3d
(green) and 4d (blue) metals.
(Bottom) reciprocal lattice of
orthorhombic perovskite
antiferromagnets, with
structural (black) and
magnetic (red) Bragg
reflections. Circles indicate
the maximal coverage of
RIXS with photons at the Cu
(green) and Ru (blue)
L2,3 -edges. (Top panel
adapted from Ref. [7], ©
American Physical Society
2011)

interested in the magnetic moment orientation, which can be extracted from the
dependence of the scattered intensity on the photon polarization without reference
to such calculations. In spherical symmetry, the scattering tensor can be expressed
in the following way [18]:

Fj (E) = σ (0) (E) εi · εo∗ + σ (1) (E) εi × εo∗ · M j


 
  ∗ 1 (10)
+ σ (2) (E) εi · M j εo · M j − εi · εo∗ ,
3

where Mj is the magnetization vector of the ion j, εi and εo are the polarization vec-
tors of the incoming and outgoing photons, and σ (0) , σ (1) , and σ (2) are proportional
to the X-ray absorption (XAS), x-ray magnetic circular dichroism (XMCD), and x-
ray magnetic linear dichroism (XMLD) tensors, respectively. Additional terms arise
1264 J. W. Lynn and B. Keimer

from the crystal field, but they tend to be small for collinear spin structures, as long
as M points along a high-symmetry direction of the crystal lattice [18].
To separate magnetic scattering from charge scattering (first term in Eq. 10),
magnetic X-ray scattering experiments can be carried out in crossed linear polariza-
tion. With the caveats mentioned above, the intensity of a magnetic Bragg reflection
of a collinear antiferromagnet at the reciprocal lattice vector g can then be written
as
2

 ig·r (1)

I = e j σj (E) εi × εo · M j ,

(11)
j

where the summation runs over the magnetic unit cell. To determine the spin
structure of a given material, one commonly uses the so-called “azimuthal scan”
where the momentum transfer g is kept fixed, and the sample is rotated such that
the orientation of M varies relative to the photon polarization vectors. In this way,
simple spin structures can be determined based on a single Bragg reflection.
Even for simple spin structures, however, it is important to keep in mind that the
spectral functions σ (E) are tensors with properties that may be strongly influenced
by the symmetry of the crystal lattice. If the site symmetry is tetragonal, for instance,
the XAS spectra for light polarized in the xy-plane and along the z-axis, σ (0) xy and
σ (0) z , are generally different – a phenomenon known as “natural linear dichroism”–
σ (1) and σ (2) are also generally anisotropic.
The deviations from spherical symmetry are particularly prominent in situations
where orbital order is present. An elementary example is the Cu2+ ion with electron
configuration 3d9 (i.e., a single hole in the d-electron shell) [18]. Materials based
on Cu2+ usually exhibit Jahn-Teller distortions that lift the degeneracy between
d-orbitals of x2 -y2 and 3z2 -r2 symmetry. The lobes of these orbitals are extended
in the xy-plane and along the z-axis, respectively. For instance, the cuprate high-
temperature superconductors exhibit a tetragonal structure with holes in the x2 -y2
orbital. In this case, the electric dipole selection rules prohibit excitation of a
2p core electron into the valence shell with z-polarized light, so that σ (0) z = 0
whereas σ (0) xy = 0. The selection rule completely changes the azimuthal scans,
as observed in resonant elastic scattering experiments on copper-oxide compounds
[19]. This example illustrates the important influence of orbital order on azimuthal
scans in magnetic X-ray scattering. Proper consideration of the crystal symmetry
is especially important for experiments performed with polarized incident light, but
without polarization analysis of the scattered beam, because magnetic and charge
scattering may then both contribute to the detected signal.
The photon energy dependence of the scattering tensor σ(E) contains a lot of
additional information, some of which can be extracted without extensive model
calculations. In particular, the large enhancement of the scattering intensity at the
absorption edges of magnetic metal atoms gives rise to the element sensitivity of
25 Magnetic Scattering 1265

magnetic X-ray scattering, which is particularly useful for multinary compounds


and for magnetic multilayers with different magnetic species. In principle, resonant
magnetic X-ray scattering is also sensitive to the valence state of metal ions, which
can be inferred from the maximum of σ (E). Resonant scattering experiments on
mixed-valent compounds have indeed been reported [20]. However, the analysis
and quantitative interpretation of such experiments require careful consideration of
the multiplets in the intermediate state.
In the discussion so far, we have not considered the spin-orbit coupling in the
valence shell, which is generally weak for 3d metal compounds. In 4f and 5f electron
systems, however, the spin-orbit coupling is so strong that it dominates the inter-
atomic exchange interactions, so that models of such compounds are based on firmly
locked spin and orbital angular momenta. In 4d and 5d electron systems, on the
other hand, the intra-atomic spin-orbit coupling turns out to be comparable to other
important energy scales including the on-site Coulomb interactions and the inter-
atomic exchange coupling. Comparative magnetic X-ray diffraction experiments at
the L2 and L3 absorption edges have recently proven to be a powerful probe of the
spin-orbit composition of the ground state wave function in such materials [46].

Dynamics

Inelastic Neutron Scattering Technique

Neutrons can also scatter inelastically, to reveal the magnetic fluctuation spectrum of
a material over wide ranges of energy (≈10−8 → 1 eV) and over the entire Brillouin
zone. Neutron scattering plays a truly unique role in that it is the only technique
that can directly determine the complete magnetic excitation spectrum, whether it
is in the form of the dispersion relations for spin wave excitations, wave-vector
and energy dependence of critical fluctuations, crystal field excitations, magnetic
excitons, or moment/valence fluctuations. In the present overview we will discuss
some of these possibilities.
As an example, consider identical spins S localized on a simple cubic lattice, with
a coupling given by -JSi ·Sj , where J is the Heisenberg exchange interaction between
neighbors separated by the distance a. The collective excitations are magnons (see
 Chap. 6, “Spin Waves”). If we have J > 0 so that the lowest energy configuration
is where the spins are parallel (a ferromagnet), then the magnon dispersion along
the edge of the cube (the [100] direction) is given by
 
Eq = 8 J S sin2 (qa/2) (12)

At each wave vector q a neutron can either create a magnon at (q, E) with a
concomitant change of momentum and loss of energy of the neutron, or conversely
1266 J. W. Lynn and B. Keimer

destroy a magnon with a gain in energy. The observed change in momentum and
energy for the neutron can then be used to map the magnon dispersion relation.
Neutron scattering is particularly well suited for such inelastic scattering studies
since neutrons typically have energies that are comparable to the energies of
interesting collective excitations in the solid, and therefore the neutron energy
changes are large and easily measured.
Additional information about the nature of the excitations can be obtained by
polarized inelastic neutron scattering techniques, which are finding increasing use.
The cross section for spin wave scattering from a simple Heisenberg ferromagnet is
given by Refs. [1, 9, 13]

 ±  2
d 2σ γ e2 k ’ S (2π )3   
= f 2
(g) nq +1/2∓1/2 δ E∓Eq δ (K∓q−g)
d 2 2mc 2 k 2 V q,g
 
× 1 + (K · η)2 ∓ 2 (P · K) · (K · η)
(13)

where nq is the Bose thermal population factor and η̂ is a unit vector in the direction
of the spins. Generally spin wave scattering is represented by the familiar raising
and lowering operators S± = Sx ± iSy , which cause a reversal of the neutron spin
when the magnon is created or destroyed. These spin-flip cross sections are denoted
by (+ −) and (− +). If the neutron polarization P is parallel to the momentum
transfer K, P||K, then the spin angular momentum is conserved (as there is no
orbital contribution in this case). In this experimental geometry, Eq. (13) shows
us that we can only create a spin wave in the (− +) configuration, which at the
same time causes the total magnetization of the sample to decrease by one unit
(1 μB for a spin-only system). Alternatively, we can destroy a spin wave only in the
(+ −) configuration, while increasing the magnetization by one unit. This gives us a
unique way to unambiguously identify the spin wave scattering, and polarized beam
techniques in general can be used to distinguish magnetic from nuclear scattering in
a manner similar to the case of Bragg scattering.
Finally, we note that the magnetic Bragg scattering is comparable in strength to
the overall magnetic inelastic scattering. However, all the Bragg scattering is located
at a single point in reciprocal space, while the inelastic scattering is distributed
throughout the three-dimensional Brillouin zone. Hence when actually making
inelastic measurements to determine the dispersion of the excitations one can only
observe a small portion of the dispersion surface at any one time, and thus the
observed inelastic scattering is typically two to three orders of magnitude less
intense than the Bragg peaks. Consequently, these are much more time-consuming
measurements, and larger samples are needed to offset the reduction in intensity.
Of course, a successful determination of the dispersion relations yields a complete
determination of the fundamental magnetic interactions in the solid.
25 Magnetic Scattering 1267

Resonant Inelastic X-ray Scattering Technique

The mechanism underlying magnetic resonant inelastic X-ray scattering (RIXS) is


analogous to the one for resonant elastic scattering discussed in Sect. “Resonant
Magnetic X-ray Diffraction Technique” and depicted in Fig. 1. A photon tuned
to a dipole-allowed transition promotes a core electron into the partially occupied
valence shell. In the intermediate state, the core-hole spin-orbit coupling induces an
electronic spin-flip, so that the reemitted photon leaves a magnetically excited state
behind. Single magnetic excitations are then observable in crossed polarization,
analogous to elastic magnetic scattering (10) [7]. In this sense, the relationship
between elastic and inelastic resonant X-ray scattering is analogous to the one
between elastic and inelastic neutron scattering. Another useful analogy is optical
Raman scattering, where single magnetic excitations at q = 0 can be activated
by the spin-orbit coupling in the intermediate state [21] which is, however,
usually much weaker than the core-hole spin-orbit coupling in RIXS. A more
common Raman scattering experiment addresses bi-magnon excitations that do
not involve an electronic spin-flip. Such experiments are also possible with RIXS
in parallel polarization geometry. As in optical Raman scattering, however, they
only determine the Brillouin-zone averaged spectrum of magnetic excitations. The
unique advantage of single-magnon RIXS is that the full magnon dispersion can be
determined even for single crystals of micrometer dimensions, or for atomically thin
films and heterostructures.
From an instrumental perspective, RIXS experiments on magnetic excitations
are challenging because the energy of the photons required to induce the atomic
dipole transition (E = 0.4–1 keV for 3d metal L-edges) largely exceeds the
typical energy of magnons in solids. A breakthrough was achieved in 2009, when
the resolving power of soft x-ray RIXS instrumentation passed the threshold of
E/ E ≈ 10,000. This enabled the first RIXS observation of high-energy magnons
in undoped layered cuprates, which exhibit an exceptionally large bandwidth of
≈300 meV [22]. Shortly thereafter, high-energy paramagnons were also observed
by RIXS in superconducting cuprates [23, 24] and in iron-based high-temperature
superconductors at the Fe L2,3 edges [25]. Kinematical constraints analogous to
those in resonant elastic scattering restrict these experiments to a fraction of the
Brillouin zone that does not include the magnetic ordering wave vectors of the
respective parent compounds. The kinematical constraints are even more severe for
RIXS experiments of bi-magnon excitations in metal oxides at the oxygen K-edge
(1 s-2p, 0.5 eV) [26].
Parallel advances in RIXS instrumentation for hard X-rays allowed the obser-
vation of single magnons in antiferromagnetically ordered iridium oxides with 5d
electron systems [27]. The larger resonance energies of the 2p-5d transition, with
correspondingly larger photon wave vectors, allow the detection of magnons over
the entire Brillouin zone. Instrumentation for RIXS at the L-absorption edges of 4d
metals and M-edges of actinides at intermediate photon energies (2.5 ≤ E ≤ 5 eV)
has only recently been developed [28].
1268 J. W. Lynn and B. Keimer

In contrast to inelastic neutron scattering, the theoretical description of RIXS


is still under development, and several open questions are actively debated in the
literature. These include the separation of spin excitations from orbital excitations
in multi-orbital systems, and from charge excitations in metallic systems. This
challenge is particularly severe in the iron pnictides, which are metals with multiple
Fermi surfaces originating from different Fe d-orbitals. A complete resolution of
this problem will likely require a transition to full polarization analysis in RIXS,
so that the different excitation channels can be separated completely. The first
experiments using RIXS polarimeters have already been reported [29]. Another
open issue is the influence of the core-hole potential in the RIXS intermediate state
of the valence electron system in metallic systems, where the core-hole lifetime may
be comparable to intrinsic timescales of the valence electrons.

Magnetic Diffraction Examples with Neutrons

As an example of magnetic powder diffraction, the scattering from a sample of


Na5/8 MnO2 is shown in Fig. 2 [30]. This material exhibits Mn3+ and Mn4+
charge stripes and vacancy ordering of the Na subsystem, which results in a rather
complicated low-temperature magnetic structure that can be determined from this
pattern. Of course, Rietveld refinements for the crystallographic structure can be
performed from the full patterns at both high and low temperatures to determine
the full crystal structure: lattice parameters, atomic positions in the unit cell, site
occupancies, etc., as well as the value of the ordered moment. The inset shows the
temperature dependence of the magnetic peak intensity, which we see from (4) is the
square of the sublattice magnetization – the order parameter of the magnetic phase
transition. Note that we can identify the magnetic scattering through its temperature
dependence, as magnetic Bragg peaks vanish above the Néel temperature along with
long range magnetic order. Note also that the magnetic intensities become weak at
high scattering angles as f (g) falls off with increasing scattering angle.
A more elegant way to identify magnetic scattering is to employ the neutron
polarization technique, particularly if the material has a crystallographic rearrange-
ment or distortion associated with the magnetic transition. It is more involved
and time consuming experimentally, but yields an unambiguous identification and
separation of magnetic and nuclear Bragg peaks. Figure 3 shows the polarized
beam results for two peaks of polycrystalline YBa2 Fe3 O8 [31]. The top section
of the figure shows the data for the P⊥g configuration. The peak on the left has
the identical intensity for both spin-flip and non-spin-flip scattering, and hence
we conclude that this scattering is purely magnetic in origin. The peak on the
right has strong intensity for (+ +), while the intensity for (− +) is smaller by
the instrumental flipping ratio. Hence this peak is a pure nuclear reflection. The
center row shows the same peaks for the P||g configuration, while the bottom
row shows the subtraction of the P⊥g spin-flip scattering from the P||g spin-flip
scattering. In this subtraction procedure, instrumental background, as well as all
nuclear scattering cross sections, cancel, isolating the magnetic scattering. We see
25 Magnetic Scattering 1269

Fig. 2 Magnetic diffraction pattern for Na5/8 MnO2 , obtained by subtracting the crystallographic
diffraction pattern obtained at 100 K, above the antiferromagnetic phase transition, from the data
at 2.5 K in the magnetic ground state. The structural scattering cancels in the subtraction if there
is no significant change when the sample magnetically orders. The inset shows the temperature
dependence of the intensity of the strongest magnetic peak, and reveals a transition temperature of
≈60 K. (Adapted from Ref. [30], © Spinger Nature 2014)

that there is magnetic intensity only for the low angle position, while no intensity
survives for the peak on the right, unambiguously establishing that the one peak is
purely magnetic and the other purely nuclear. These data also demonstrate that all
three components of the angular momentum contribute to the magnetic scattering.
This simple example demonstrates how the technique works; obviously it plays a
more critical role in cases where it is not clear from other means what is the origin
of the peaks, such as in regimes where the magnetic and nuclear peaks overlap, or
in situations where the magnetic transition is accompanied by a structural distortion
where the structural peaks change significantly in intensity.
When investigating the magnetic structures of new materials, it is generally best
to first carry out powder diffraction experiments to establish the basic properties
of the magnetic structure, assuming of course that the ordered moment is large
enough to observe the magnetic Bragg peaks. Once the basics are established,
on the other hand, measurements on a single crystal can provide much higher
1270 J. W. Lynn and B. Keimer

Fig. 3 Polarized neutron diffraction on polycrystalline YBa2 Fe3 O8 . The top portion of the figure
is for P⊥g, where the open circles show the non-spin-flip scattering and the filled circles are in the
spin-flip configuration. The low angle peak has equal intensity for both cross sections, and thus is
identified as a pure magnetic reflection, while the ratio of the (+ +) to (− +) scattering for the high
angle peak is just the instrumental flipping ratio. Hence this is a pure nuclear reflection. The center
portion of the figure is for P||g, and the bottom portion is the subtraction of the spin-flip data for
the P⊥g configuration from the spin-flip data for P||g. Note that in the subtraction procedure all
background and nuclear cross sections cancel, thereby isolating the magnetic scattering. (Reprinted
by permission from Ref. [31], © American Physical Society 1992)
25 Magnetic Scattering 1271

Fig. 4 Neutron diffraction intensity map observed in the (h, k, 0) scattering plane of a single
crystal of the multiferroic Co3 TeO6 . The temperature is 22 K, just below the antiferromagnetic
phase transition at TN = 26 K. The nuclear Bragg peaks at integer positions are accompanied by
four satellite magnetic reflections, indicating the development of incommensurate (ICM) magnetic
order. Note that the ordering wave vector is incommensurate in both h and k. No energy analyzer
was used for these measurements so that the data are energy-integrated, and there is clear diffuse
scattering surrounding the ICM peaks at this temperature originating from magnetic excitations.
(Adapted from Ref. [32], © American Physical Society 2012)

quality and more detailed information about the magnetic properties. Figure 4
shows a map of the scattering intensity in the (h,k,0) scattering plane at 22 K for
a single crystal of the multiferroic Co3 TeO6 , which orders antiferromagnetically at
26 K [32]. The crystal structure is monoclinic, and we see four satellite magnetic
peaks around each (integer) structural peak, indicating that the initial magnetic
structure is incommensurate in both the h and k (and l as well, it turns out [33])
directions. With further decrease of temperature, a series of additional transitions
are observed, details of which would be difficult to determine with a powder. At
lower temperatures, separate commensurate peaks develop, then there is a lock-
in transition along k that includes a ferroelectric order parameter, and then finally
a transition into the ground state with both commensurate magnetic order and
incommensurate order along h, k, and l. [33, 34].
The magnetic superconductor ErNi2 B2 C goes superconducting at TC = 11 K,
and then develops incommensurate antiferromagnetic order below TM = 6 K as
shown in Fig. 5 [35]. The wave vector for the ordering is (h,0,0) with h ≈ 0.55, with
the spin direction transverse, along (0,y,0). Initially, the magnetic order exhibits a
1272 J. W. Lynn and B. Keimer

Fig. 5 (Top) Unpolarized neutron diffraction measurements along the (h,0,0) direction at 1.3 K,
2.4 K, and 4.58 K of a single crystal of ErNi2 B2 C. At 10 K no peaks are observed in this wave
vector range. The data have been offset along the intensity axis for clarity. Above the weak
ferromagnetic transition at 2.3 K the fundamental incommensurate peak is observed at h = 0.55,
along with higher odd-order harmonics. Below the ferromagnetic transition a new set of even-
order harmonics develops, indicated by the arrows. (Bottom) Schematic of the initial transversely
polarized spin-density-wave, and ground state square-wave. (Adapted from Ref. [35], © American
Physical Society 2001)
25 Magnetic Scattering 1273

simple sinusoidal spin-density-wave (SDW) that is transversely polarized, as shown


in the bottom of the figure. As the amplitude of the SDW increases, third-, fifth-, and
higher-order wave vector peaks develop as the wave squares up. This is the expected
behavior since for localized moments entropy mandates that a simple spin density
wave cannot be the ground state magnetic structure.
For any SDW structure, only odd-order peaks will have nonzero intensity due to
time-reversal symmetry, because on average the net magnetization is zero. Below
2.3 K we see that a new set of even-order peaks is found along the (h,0,0)
direction of ErNi2 B2 C. One possibility is that the even-order peaks are due to
a structural distortion, a charge-density wave (CDW) that follows the SDW due
to a magnetoelastic interaction. Hence the even-order peaks would be structural
peaks and the odd-order peaks magnetic. In the present material, however, a net
magnetization develops in the superconducting state in the magnetic ground state,
so that the even-order peaks could be structural, magnetic, or both. To establish
the nature of these peaks unambiguously, polarized neutron diffraction was used
as shown in Fig. 6. The data were measured in the (h,0,l) scattering plane, with k
then perpendicular to the scattering plane. For P||g the spins are perpendicular to
the scattering plane and hence perpendicular to P and then the magnetic scattering
is all spin-flip. Note that the polarization dependence of the cross sections is
quite different than the YBa2 Fe3 O8 example above, emphasizing that the spin-flip
and non-spin-flip magnetic cross sections depend on the details of the magnetic
structure. The structural scattering is always non-spin-flip. The data show that both
odd-order and even-order are purely magnetic in this system.
For antiferromagnets there is no net magnetization produced by the magnetic
ordering. When the sublattice magnetizations are not compensated and there is a
net magnetization, on the other hand, the superconductivity must respond to and try
to screen this magnetization. If the internally generated field is below HC1 then the
supercurrents will exactly compensate the net magnetization and the total field will
be zero. If the field exceeds HC2 then the superconductivity will be extinguished
as happens in materials such as ErRh4 B4 and HoMo6 S8 [36]. Between these two
cases, vortices are expected to be spontaneously generated, and this possibility
can be investigated with SANS. Figure 7 shows SANS data from a single crystal
of ErNi2 B2 C [37]. The inset presents the image on the two-dimensional SANS
detector, where K = 0 is in the center. We see the expected hexagonal pattern
of scattering from the vortex lattice. Below the ferromagnetic transition additional
vortices spontaneously form due to the internally generated magnetic field, which
adds to the applied field. To accommodate the additional vortices they rearrange
themselves with a smaller lattice parameter for the vortex lattice, which is reflected
by the peak of the vortex scattering moving to larger K [38].
The above examples demonstrate scattering from long-range magnetic order
where the magnetic diffraction consists of resolution-limited Bragg peaks. But that
is not always the case, and some of the best examples occur where competing
magnetic interactions lead to frustration and suppress the order to lower temper-
atures or prevent it completely. Arguably the best example of a frustrated lattice
1274 J. W. Lynn and B. Keimer

Fig. 6 Polarized neutron diffraction measurements on a single crystal of ErNi2 B2 C showing both
the odd-order (5th) and even-order (16th) harmonics for the P||g configuration. The solid circles
(−, +) and solid triangles (+, −) are spin-flip scattering, while the open circles (+, +) and open
triangles (−, −) are non-spin-flip scattering. The data demonstrate that both types of reflections
are magnetic in origin, with the moment direction along the b axis. (Adapted from Ref. [35],
© American Physical Society 2001)

occurs in the cubic rare-earth (R) pyrochlore (R2 Ti2 O7 ) systems where the R ions
occupy corner-sharing tetrahedra [39]. For R = Ho, Dy, for example, the single-
ion anisotropy restricts the moments to point along diagonal (111) directions, along
lines that intersect the center of each tetrahedron. The ground state turns out to be
with two of the moments pointing into each tetrahedron and two pointing out. But
you do not know which two are in and which two are out, exactly like the hydrogen
bonding in ice where two H move into the oxygen in the center of the tetrahedron
and bond and two move out, resulting in a macroscopic degeneracy that violates the
third law of thermodynamics. The first measurement of the ground state correlations
was carried out for Ho2 Ti2 O7 , where the observed scattering from the correlated
moments agreed quite well with simulations [40]. An interesting simplification
occurs for a field applied along the (111) direction, which isolates the layers and
forms two-dimensional “kagomé spin-ice.” The scattering for this case is shown
25 Magnetic Scattering 1275

Fig. 7 Radially averaged small angle neutron scattering intensity of the vortex scattering in
ErNi2 B2 C vs wave vector K at 85 mT, above and below the weak ferromagnetic transition.
The shift in the peak position demonstrates that additional vortices spontaneously form as the
macroscopic magnetization develops at low temperatures. The temperature dependence shows that
this spontaneous vortex formation is directly related to the weak ferromagnetic transition. The inset
shows vortex Bragg peaks on the two-dimensional SANS detector; K = 0 is in the center. (Adapted
from Ref. [37])

in Fig. 8 for Dy2 Ti2 O7 , which shows the broad distributions of diffuse magnetic
scattering that are in excellent agreement with Monte Carlo simulations [41].
The ground state properties are not the only remarkable property of spin-ice,
as the magnetic excitations are equally fascinating. Theory showed that these
excitations, which simply consist of flipping one of the spins in a tetrahedron so
that you have three pointing out and one pointing in (and in the adjacent tetrahedron
three point in and one out), correspond to the creation of a magnetic monopole and
anti-monopole [42]. The subsequent motion of these particles is governed by the
Coulomb Hamiltonian for magnetic charges, and this scenario was subsequently
confirmed by neutron scattering measurements [43, 41, 44].
Advances in thin film deposition methods have facilitated the synthesis of com-
plex heterostructures with atomic layer accuracy, which has enabled investigators
1276 J. W. Lynn and B. Keimer

Fig. 8 (a) Neutron measurements of the diffuse magnetic scattering in the kagomé spin-ice
compound Dy2 Ti2 O7 at T = 0.43 K and B = 0.5 T. The sharp structural Bragg peaks, such as
(2,-2,0), are contained within one pixel and have been removed from the plot. (b) Monte Carlo
simulations of the expected scattering in this kagomé spin-ice state. The overall features are in
excellent agreement with the data. (Adapted from Ref. [41], © The Physical Society of Japan
2009)

to control the magnetic properties by tailoring the exchange interactions within


and between layers. These capabilities combined with advances in experimental
reflectometry techniques have made neutron scattering an essential tool to elucidate
the atomic depth profile and magnetization density of thin films and multilayers.
An interesting example is the multilayer oxide heterostructure consisting of the
(approximately cubic) antiferromagnets LaMnO3 and SrMnO3 , grown on a SrTiO3
substrate. The structural indices of refraction for these two materials are almost
identical, rendering nuclear scattering from the mutilayer superstructure practically
invisible. Occasionally an extra layer of LaMnO3 was deposited to dope the
interface, which produced an effective composition of La0.44 Sr0.56 MnO3 , which is
in the ferromagnetic regime. Figure 9 shows the non-spin-flip polarized neutron
reflectivity data in the two polarization states, R++ and R− − , that are sensitive to
the ferromagnetism. The resulting magnetic depth profile reveals that the magnetic
modulation is quite large, varying from 0.7 to 2.2 μB , and that its period corresponds
precisely to the LMO superlattice structure [45]. High angle diffraction data on
the epitaxial multilayer confirmed the canted modulated spin structure of the
superlattice.

Magnetic Diffraction Examples with X-rays

As an example of resonant magnetic X-ray scattering, we first highlight experiments


on the antiferromagnet Sr2 IrO4 with hard X-rays tuned to the Ir L2 ,3 edges [46]. The
crystal structure of Sr2 IrO4 is composed of IrO2 square lattices, closely similar to
La2 CuO4 , the parent compound of a prominent family of high-temperature super-
conductors. Prior to the X-ray experiments, magnetic susceptibility measurements
had suggested antiferromagnetic order with a Néel temperature of 240 K, but
neutron diffraction experiments had proven difficult because of the large neutron
absorption cross section of Ir, and because large single crystals could not be
25 Magnetic Scattering 1277

Fig. 9 (a) Non-spin-flip polarized neutron reflectivity data R++ (red) and R− − (blue) on a
LaMnO3 /SrMnO3 multilayer, measured in a 675 mT field at 120 K. The inset shows a schematic
of the superlattice. (b) Magnetic depth profile determined by the fit to the data. Location of
the LaMnO3 (pink) and SrMnO3 (green) layers are shown. (c) Spin-flip intensity, showing the
antiferromagnetic peak and satellite peak. Inset shows the non-spin-flip scattering in the same
range. (Adapted from Ref. [45], © American Physical Society 2011)
1278 J. W. Lynn and B. Keimer

grown. The hard X-ray data on a crystal of sub-millimeter dimensions show


multiple magnetic Bragg reflections that can be analyzed by refining the Bragg
intensities according to Eq. (11) in a manner entirely analogous to magnetic neutron
diffraction. The analysis revealed a canted antiferromagnetic structure in the IrO2
planes, with alternating stacking in the direction perpendicular to the planes.
The photon energy dependence of the resonant magnetic X-ray scattering cross
section yields additional information about the magnetic ground state of Sr2 IrO4 that
would be difficult to obtain with neutron diffraction, even under ideal conditions.
The Ir valence electrons occupy 5d orbitals of xy, xz, and yz symmetry. For materials
with 3d valence electrons, the crystal field lifts the degeneracy between these
orbitals and quenches the orbital magnetization. In the 5d electron shell, however,
the strong intra-atomic spin-orbit coupling can generate complex admixtures of
these orbitals in the ground-state wave function, which correspond to a nonzero
orbital magnetic moment. This, in turn, affects the matrix elements for the photon-
induced transitions from the spin-orbit split 2p shell into the 5d shell such that the
diffraction intensities at the L2 and L3 edges (2p1/2 -5d and 2p3/2 -5d, respectively)
can become different. The strong disparity of the diffraction intensities observed
experimentally (Fig. 10) [46] indicates that the orbital magnetization is largely
unquenched, and that the spin and orbital components of the magnetic order
parameter in Sr2 IrO4 are of comparable magnitude. Similar observations have been
made for other iridates. Models of magnetism in the iridates are therefore commonly
expressed in terms of the total angular momentum, Jeff = S + L. For Sr2 IrO4 ,
Jeff = ½ in the ground state.
The large resonant scattering cross section, combined with the high photon flux
at synchrotron beamlines and the focusing capability of advanced x-ray instrumen-

Fig. 10 Photon energy dependence of the (1, 0, 22) magnetic Bragg reflection at the L3 -(left) and
L2 -edges (right) of Sr2 IrO4 . The black lines show the x-ray absorption spectra for comparison.
(Reprinted with permission from Ref. [46], © American Association for the Advancement of
Science 2009)
25 Magnetic Scattering 1279

Fig. 11 Map of the resonant elastic X-ray scattering intensity at the (0, 0, 1) magnetic Bragg
reflection of La0.96 Sr2.04 Mn2 O7 at the Mn L3 -edge. The data indicate domains where the Mn spins
point in different directions in the MnO2 layers. (Reprinted with permission from Ref. [47], ©
American Physical Society 2013)

tation, allow magnetic X-ray scattering experiments with beam dimensions well
below typical magnetic domain sizes. Figure 11 provides an example of such an
experiment on the layered antiferromagnet La0.96 Sr2.04 Mn2 O7 , which comprises
alternately stacked sheets of ferromagnetically aligned Mn spins [47]. The (001)
1280 J. W. Lynn and B. Keimer

magnetic Bragg reflection of this spin array can be reached with photons tuned to
the Mn L3 -edge. The data shown in Fig. 11 were taken with a beam of 300 nm
diameter. They reveal that the diffracted intensity varies on a characteristic length
scale of several microns. A detailed analysis shows that the intensity variation results
from domains with different spin directions, which diffract photons with different
scattering amplitude due to the photon polarization dependence of the scattering
cross section (Eq. 11). In another study, domains with different helicities in a spiral
magnet were imaged by resonant diffraction with circularly polarized X-rays. The
spatial resolution and imaging capabilities of magnetic X-ray scattering methods
are expected to develop rapidly with the advent of coherent X-ray beams at fourth-
generation synchrotron sources.
In analogy to neutron reflectometry, polarized magnetic X-ray reflectometry has
recently developed into a powerful, element-sensitive probe of complex oxide thin
films, heterostructures, and superlattices. As an example, we discuss resonant X-
ray diffraction data on RNiO3 -based films and superlattices (where R denotes a
lanthanide atom). RNiO3 perovskites exhibit a Mott metal-insulator transition as
a function of the radius of the R cation, which modulates the Ni-O-Ni bond angle.
Recent work has shown that the metal-insulator transition can also be controlled
by epitaxial strain and by spatial confinement of the conduction electron system.
Antiferromagnetism with ordering vector g = (¼, ¼, ¼) develops in the Mott-
insulating phase. Figure 12 (top) shows azimuthal scans at the corresponding
magnetic Bragg reflection taken with photons tuned to the Ni L3 -edge [48]. The
data analysis demonstrates that the magnetic order is noncollinear, with Ni spins
forming a spiral propagating along the (111) direction of the perovskite unit cell.
The polarization plane of the spiral can be controlled by epitaxial strain.
Figure 12 (bottom) shows a contour map of the resonant scattering intensity from
a LaNiO3 -LaAlO3 superlattice as a function of the azimuthal angle and the momen-
tum transfer perpendicular to the superlattice plane [49]. Strong modifications of
the azimuthal-angle dependence of the intensity occur particularly under grazing-
incidence or grazing-exit conditions, where the incident or scattered beams are
strongly refracted at the external and internal interfaces of the superlattice. These
data illustrate the possibly important influence of dynamical effects in resonant
soft X-ray diffraction from thin-film structures, which go beyond the kinematic
approximation usually employed in the analysis of such data.
Very recently, X-ray free-electron lasers have enabled time-resolved resonant
magnetic diffraction experiments capable of imaging the real-time dynamics of
magnetic order under nonequilibrium conditions. As an illustration of this emerging
capability, Fig. 13 shows the time evolution of the g = (¼, ¼, ¼) antiferromagnetic
Bragg peak of a NdNiO3 film following a THz pump pulse exciting an infrared-
active phonon mode of the LaAlO3 substrate [50]. As the phonon propagates from
the substrate through the film, it obliterates the antiferromagnetic order in its wake
on a picosecond timescale. The mechanism underlying this “non-thermal melting”
phenomenon may involve transient distortions of the NiO6 octahedra, which weaken
the magnetic exchange interactions between Ni spins.
25 Magnetic Scattering 1281

Fig. 12 (Top left) Ni L3 -edge scans through the (¼, ¼, ¼) magnetic Bragg reflection of LaNiO3 -
LaAlO3 superlattices with different numbers of consecutive unit cells. The absence of the magnetic
Bragg peak in superlattices with three or more LaNiO3 layers indicates that the magnetic order in
the 2×2 superlattice is induced by spatial confinement of the conduction electrons. (Reprinted
with permission from Ref. [48]). (Top right) azimuthal angle dependence of the (¼, ¼, ¼)
magnetic Bragg peak of nickelate thin films and superlattices with simulations that rule out
collinear (CM) and favor noncollinear (NCM) magnetism. (Bottom) simulated contour map of
the scattering intensity of 2×2 LaNiO3 -LaAlO3 superlattice as functions of azimuthal angle
and momentum transfer perpendicular to the superlattice plane, demonstrating the importance of
dynamical diffraction effects. (Reprinted with permission from Ref. [49], © American Physical
Society 2016)
1282 J. W. Lynn and B. Keimer

Fig. 13 (a) Schematic illustration of the demagnetization process of a NdNiO3 film triggered
by a coherently excited phonon of the LaAlO3 substrate. (b) Depth profile of the (¼, ¼, ¼)
resonant magnetic Bragg peak intensity at different time delays between the phonon pump pulse
and the resonant X-ray diffraction probe measurement. (Reprinted with permission from Ref. [50],
©Springer Nature 2015)

Spin Dynamics with Neutrons

There are many types of magnetic excitations and fluctuations that can be measured
with neutron scattering techniques, such as magnons, spinons, critical fluctuations,
crystal field excitations, magnetic excitons, and moment/valence fluctuations. We
start with classic magnons in an isotropic ferromagnet, where the excitations are
gapless and the dispersion relation is given by Eq. (12). Figure 14 (left) shows a
measurement for La0.67 Ca0.33 MnO3 , which is a colossal magnetoresistive (CMR)
material [51]. The data reveal two magnon peaks at a given wave vector, one in
energy gain where the neutron destroys a magnon and gains energy, and one in
energy loss where a magnon is created. This is a small q (long wavelength) excita-
tion, and in fact this sample is polycrystalline rather than single crystal, and the data
were collected around the (0,0,0) reciprocal lattice position. Such measurements are
restricted in wave vector and energy, and are only viable for isotropic ferromagnets;
otherwise the excitations fall outside the accessible experimental window dictated
by momentum and energy conservation. If there is a question of whether these
excitations are magnons or phonons, the polarized beam technique can be employed
as shown in Fig. 14 (right) for the prototypical isotropic ferromagnet amorphous
Fe86 B14 [52]. These data were taken with the neutron polarization P parallel to the
momentum transfer K (P||K), Eq. (13) with q = K. In this configuration magnons
require the neutron spin direction to reverse (spin-flip), while phonons can only be
observed in the non-spin-flip configuration. For magnons we should be able to create
25 Magnetic Scattering 1283

La0.67Ca0.33MnO3
1400 Horizontal Field Q=0.09 T=500K
–1 200 K
Q = 0.07Å
1200
300
1000 (+–)
(–+)
Counts/7 min.

Counts/30 min
800
200

600

400 100

200

0 0
–1.0 –0.5 0.0 0.5 1.0
–0.8 –0.6 –0.4 –0.2 0 0.2 0.4 0.6 0.8
Energy (meV) Energy (meV)

Fig. 14 Spin waves in isotropic ferromagnets. (Left) energy scan at a wave vector q of 0.07 Å−1
for La0.7 Ca0.3 MnO3 , (Reprinted with permission from Ref. [51], © American Physical Society
1996) showing the spin waves in energy gain (E < 0) and energy loss (E > 0). (Right) polarized
beam energy scan on the Fe86 B14 amorphous ferromagnet at a fixed wave vector of 0.09 Å−1 , with
the neutron polarization parallel to q. In this configuration spin angular momentum is conserved,
and the neutron can only create an excitation (E > 0) if its moment is initially antiparallel to the
magnetization, and can only destroy a spin wave (E < 0) when its moment is parallel. (Reprinted
with permission from Ref. [52], © American Institute of Physics 1996)

a spin wave only in the (− +) configuration where the incident neutron moment is
antiparallel to the magnetization; the scattered neutron moment is then parallel to
the magnetization direction, and the magnetization is decreased by one unit by the
creation of the magnon. On the energy gain side the process is reversed and we
destroy a magnon only in the (+ −) configuration. This is precisely what we see
in the data; for the (− +) configuration the spin waves can only be observed for
neutron energy loss scattering (E > 0), while for the (+ −) configuration spin waves
can only be observed in neutron energy gain (E < 0). This behavior of the scattering
uniquely identifies these excitations as magnons.
Expanding the sine in Eq. (12) we see that the small-q dispersion relation can
be written as Esw = D(T)q2 , where D is the spin wave “stiffness” constant. The
general form of the spin wave dispersion relation is the same for all isotropic
ferromagnets, a requirement of the (assumed) perfect rotational symmetry of the
magnetic system, while the numerical value of D depends on the details of the
magnetic interactions and the nature of the magnetism. The small-q dispersion
relation can be readily measured, as shown in Fig. 15 (left) for a single crystal of
La0.85 Sr0.15 MnO3 , and D(T) obtained [53]. The effect of temperature is to soften
the average exchange interaction as the magnetization decreases, and hence the
magnons renormalize to lower energies with increasing temperature as also shown
Fig. 15. With single crystals the dispersion curves can be determined in different
directions and throughout the Brillouin zone, as shown in Fig. 15 (right) for a
1284 J. W. Lynn and B. Keimer

La0.85Sr 0.15MnO 3
5

4 T = 10 K
Energy [meV]

T = 220 K
3

(a)
0
0 0.50 0.1 0.15 0.2 0.25 0.3 0.35
q (0, 0, ζ)
100
(b)
80
D010
D [meV Å2]

60 D 001

40
_ k T 5/2
20 Fit to D0{1 - ct. l2 ( ____
B
) ζ(5/2) }
4πD 0
Fit to a power law
0
0 50 100 150 200 250
Temperature [K]

Fig. 15 (Left) (a) Low energy spin wave dispersion relations at two different temperatures. The
dispersion relation follows a quadratic dependence expected for a ferromagnet, which defines the
spin stiffness D, and no significant gap in the excitation spectrum is observed indicating an isotropic
system. D(T) is shown in (b), which follows a power law behavior as the Curie temperature
is approached. (Reprinted with permission from Ref. [53], © American Physical Society 1998).
(Right) spin wave dispersion relations for a series of colossal magnetoresistive perovskite oxides.
(Reprinted with permission from Ref. [54], © American Physical Society 2006)

number of perovskite CMR systems [54]. Such measurements enable a detailed


determintion of all exchange interactions, rather than just the long wavelength
(average) behavior. Any gap(s) in the excitation spectrum can also be directly
measured.
In addition to the magnon energies, the lifetimes of the excitations can also be
determined by extracting the intrinsic widths of the excitations, both in the ground
state far below the ordering temperature, and as a function of temperature. An
example of the linewidths in the ground state are shown for La0.85 Sr0.15 MnO3
in Fig. 16 [53]. In the simplest localized-spin model, negligible intrinsic spin
wave linewidths would be expected at low temperatures, while we see here that
the observed linewidths are substantial at all measured wave vectors and highly
anisotropic, indicating that an itinerant electron type of model is a more appropriate
description for this system. In particular, the linewidths become very large at large
wave vectors. These substantial linewidths are easy to measure with conventional
instrumentation. Insulating magnets, on the other hand, generally have much smaller
linewidths and require much higher instrumental resolution to measure. Figure 16
25 Magnetic Scattering 1285

Fig. 16 (Left) intrinsic spin wave linewidths for the ground state magnetic excitations in
La0.85 Sr0.15 MnO3 . The linewidths are quite anisotropic, and are small at small wave vectors but
become very large at large q (Reprinted with permission from Ref. [53], © American Physical
Society 1998). (Right) magnon linewidths as a function of temperature for a series of q’s in the
insulating antiferromagnet Rb2 MnF4 , measured using the high resolution spin-echo triple-axis
technique. (Reprinted with permission from Ref. [55], © American Physical Society 2006). The
solid curves are calculations using spin wave theory

(right) shows the measured linewidths for the prototype insulating antiferromagnet
Rb2 MnF4 [55]. Here the spin-echo triple-axis technique has been employed, which
has extraordinarily good (μeV) resolution. The theoretically calculated linewidths
from spin-wave theory are shown by the solid curves at a series of temperatures, and
are in quantitative agreement with the data.
One area where neutron scattering has played an essential role is elucidating
the spin dynamics of the high temperature superconductors, first for the copper
oxide systems [56] and more recently for the iron-based superconductors [57]. The
magnetic excitations in these classes of materials extend to quite high energies –
as high as ≈0.5 eV – making the measurements particularly challenging since the
magnetic form factor requires that the magnitude of K must be kept small, neces-
sitating quite high incident energy neutrons. These requirements are well matched
to the time-of-flight capabilities of spallation neutron facilities where high energy
neutrons are plentiful. To illustrate the basic technique, consider the excitations from
BaFe2 As2 , which is one of the antiferromagnetic “parent” materials of the iron-
based superconductors. The antiferromagnetic ordering temperature TN = 138 K,
which corresponds to a thermal energy of just ≈12 meV (1 meV ➔ 11.605 K). Yet
we see from Fig. 17 that the magnons extend up to 200 meV, an order-of-magnitude
higher energies than the ordering temperature represents, indicating that the system
has a low-dimensional character [58]. The in-plane dispersion relations are also
quite anisotropic, even though the orthorhombic distortion away from tetragonal
symmetry (that accompanies the magnetic order) is small. Another very interesting
aspect of the magnetic excitations is that they have quite large linewidths at high
energies, indicating that the magnetic electrons are itinerant in nature. Indeed, the
1286 J. W. Lynn and B. Keimer

Fig. 17 (Left) constant-energy cuts of the magnetic excitations in BaFe2 As2 at a series of
energies. The solid curves are the fits to a spin wave model. (Right) spin wave dispersion along the
(1, K) direction as determined by energy and Q cuts of the raw data. The solid curve is a Heisenberg
model calculation using anisotropic exchange couplings. The dotted line is a Heisenberg model
calculation assuming isotropic exchange coupling. (Adapted from Ref. [58], © American Physical
Society 2011)

iron d-bands where the magnetism originates cross the Fermi energy – the definition
of itineracy.
Our final neutron example concerns the spin dynamics of one-dimensional (1D)
magnets, which (together with 2D magnets) have played a special role in developing
a fundamental understanding of quantum magnetic systems. This is because they
are more tractable theoretically and therefore enable a deeper comparison with
experiment. They also entail the emergence of new types of cooperative states and
their associated excitations. A classical case is the spin one-half antiferromagnet
chain, where quantum effects are maximal, represented by materials such as KCuF3
[59] and CuSO4 ·5D2 O [60], which have enjoyed a long and interesting history of
investigations. The ground state turns out to be an entangled macroscopic singlet,
but where the two-spin correlation function decays only algebraically, rendering
long lengths of the chain to be correlated antiferromagnetically. The fundamental
excitations of such a 1D system are spinons in these (isolated) spin chains, which
can be considered to a first approximation as moving domain walls. Measurements
of the dynamic structure factor for CuSO4 ·5D2 O are shown in Fig. 18 [60].
Spinons carry fractional spin, and hence these fractionalized excitations can only
be created in pairs in the scattering process. Thus the lower energy part of the
spectrum corresponds to two-spinon excitations and has the appearance of a simple
antiferromagnetic spin wave dispersion relation. However, only 71% of the spectral
weight is contained in this two-spinon component, with essentially all the remainder
being accounted for by the four-spinon contribution. Precise calculations of the
25 Magnetic Scattering 1287

0.1

w
0.0 0.0

Fig. 18 Intensity color maps of the experimental inelastic neutron scattering spectrum measured
along the Cu chain in CuSO4 ·5D2 O are shown in the left, compared with the theoretical two-
and four-spinon dynamic structure factor. (Reprinted with permission from Ref. [60], © Springer
Nature 2013)

dynamic structure factor for two-spinon and four-spinon scattering are also shown
in Fig. 18, which account for essentially the entire measured spectral weight, and
are in excellent agreement with the measurements [60].

Spin Dynamics with RIXS

The set of materials investigated by high-resolution RIXS is thus far mostly limited
to magnets with characteristic exchange interactions of the order of 100 meV.
A milestone was set by early experiments on La2 CuO4 , the antiferromagnetic,
Mott-insulating end member of a family of high-temperature superconductors,
which exhibits an exceptionally large magnon bandwidth of ≈300 meV. A RIXS
spectrometer with energy resolution of E ≈ 100 meV proved to be capable of
separating these excitations from the elastic line over a substantial fraction of the
Brillouin zone (Fig. 19) [22, 23]. Comparison with prior inelastic neutron scattering
data on the same materials demonstrated that the RIXS excitation features indeed
originate from single antiferromagnetic magnons.
RIXS experiments have also revealed the persistence of high-energy paramagnon
excitations in highly doped, superconducting cuprates. Based on the polarization
dependence of the scattering cross section at specific scattering geometries, they can
be separated from charge excitations, as shown in Fig. 20 for YBa2 Cu3 O6 + x [29].
The measurements are complementary to inelastic neutron scattering experiments,
which have much higher energy resolution and can therefore access spin excitations
with energies from 1 to 100 meV, comparable to the superconducting energy
gap. The RIXS measurements, on the other hand, are more sensitive to high-
energy excitations, which can also be investigated with high energy neutrons from
spallation sources. The photon energy dependence of the RIXS intensity yields
1288 J. W. Lynn and B. Keimer

(b) (c) 400


XAS LCO Neutrons
Norm. intensity (arb. u.)

20 Cu L3 T = 15 k 300

Energy (meV)
q = +1.885 RIXS
//
B 200
930 932
Photon energy (eV)
10
D LCO 100
C A T = 15 k
0
0
1.0 0.5 0.0 –3 –2 –1 0 1 2 3
Energy loss (eV) q
//

Fig. 19 (Left) RIXS profile of La2 CuO4 taken at the Cu L3 edge with ≈100 meV energy
resolution. The spectrum can be decomposed into elastic (A), single magnon (B), multiple magnon
(C), and optical phonon (D) components. The inset shows the X-ray absorption spectrum near
the Cu L3 edge, the arrow marks the energy of the incident photons. (Right) single magnon
dispersion determined by RIXS (blue dots), compared to inelastic neutron scattering data (dashed
line). (Reprinted with permission from Ref. [23], © American Physical Society 2010)

additional insight into the nature of these excitations. Whereas the spin excitation
energy is independent of photon energy, as expected for collective modes, the
spectral weight of the charge excitations shifts upon detuning the photon energy
away from the L-edge resonance, signaling a broad excitation continuum. This
supports models that treat collective spin excitations as mediators of unconventional
superconductivity.
Hard X-ray RIXS experiments on the layered iridates have revealed magnon
dispersions remarkably similar to those of the cuprates – a finding that has fueled
predictions of unconventional superconductivity in the iridates. In addition to the
usual low-energy magnon branches emanating from the antiferromagnetic Bragg
reflections, these experiments have also revealed weakly dispersive “spin-orbit
exciton” modes corresponding to spin excitations from the Jeff = ½ ground state
into the Jeff = 3/2 excited state (Fig. 21) [27]. Since the dispersion of these modes is
controlled by the combination of the intra-atomic spin-orbit coupling, the crystalline
electric field, and the inter-atomic exchange interactions, RIXS experiments are an
incisive probe of the low-energy electronic structure of these materials.
Finally, to illustrate the diversity of inelastic X-ray scattering methods applied
to magnetism, we highlight results of an X-ray emission spectroscopy study of iron
arsenide superconductors of composition Ca1-x Rx Fe2 As2 (where R = rare earth)
[61]. The goal of this experiment was to elucidate the origin of a pressure-induced
structural phase transition from an antiferromagnetic to a nonmagnetic state that is
associated with a large volume reduction [62, 63]. To measure the local magnetic
moment of the Fe ions independent of any inter-atomic correlations, X-ray photons
were tuned to the Fe K-absorption edge (1 s-3d), and the spectrum of emitted X-rays
was monitored around the dipole-active Kβ emission line (2p-1 s). A local moment
on the Fe site induces a splitting of this line (inset of Fig. 22) whose size depends
25 Magnetic Scattering 1289

Fig. 20 Photon energy dependence of the RIXS intensity (a, b) for undoped antiferromagnetic
YBa2 Cu3 O6.1 and (c, d) superconducting Ca-substituted YBa2 Cu3 O7 in polarization geometries
that predominantly select spin (a, c) and charge (b, d) excitations. The horizontal dashed lines
highlight the energy independence of the magnetic peak position, while the dashed green line is a
guide to the eye underlining the fluorescence behavior of the continuum of charge excitations from
the doped holes. (Reprinted with permission from Ref. [29], © American Physical Society 2015)

on the moment amplitude. These experiments led to the discovery of a pressure-


induced spin-state transition from a high-spin to a low-spin configuration of the Fe
atoms. The lower volume of the low-spin Fe atoms explains the volume collapse in
the nonmagnetic phase at high pressures.

Facilities and Online Information

A list of current neutron scattering facilities around the world can be found at
(http://en.wikipedia.org/wiki/Neutron_research_facility). Numerical values of the
free-ion magnetic form factors for neutrons can be obtained at https://www.ill.eu/
sites/ccsl/ffacts/ffachtml.html. Values of the coherent nuclear scattering amplitudes
and other nuclear cross sections can be found at http://www.ncnr.nist.gov/resources/
n-lengths/.
1290 J. W. Lynn and B. Keimer

Fig. 21 (Top) the spin-orbital level scheme of Sr2 IrO4 . The spin-orbit coupling λ splits the
d-electron manifold into Jeff = 1/2 and 3/2 multiplets. The crystal field lifts the degeneracy
of the Jeff = 3/2 multiplet. Orange (blue) colors in the images of the orbitals represent spin up
(down) projections. (Bottom) dispersion of magnons and spin-orbit excitons (marked with QP for
“quasiparticle”) extracted from RIXS data at the Ir L-edge. (Reprinted with permission from Ref.
[27], © Springer Nature 2014)

A list of current X-ray scattering facilities can be found at (http://en.wikipedia.


org/wiki/List_of_synchrotron_radiation_facilities).
Values for characteristic X-ray energies and a guide to the literature on X-ray
form factors can be found at http://xdb.lbl.gov/.
25 Magnetic Scattering 1291

Fig. 22 Fe Kβ emission spectra of Ca1-x Rx Fe2 As2 with R = Nd, and difference spectrum with
FeCrAs where Fe is in a nonmagnetic spin-0 state. The difference spectrum indicates a splitting of
the emission line due to a local magnetic moment on the Fe site (inset). (Reprinted with permission
from Ref. [61], © American Physical Society 2013)

Energy Units: Traditionally magnetic excitations are quoted in units of meV but
sometimes authors use THz, particularly for phonons in older literature. Raman
and IR experimenters often use cm−1 . 1 meV ➔ 0.24180 THz ➔ 8.0655 cm−1
➔ 11.605 K.
For a wavelength λ = 1.54 Å the photon energy is 8.05 keV, the electron energy
is 63.4 eV, and for a neutron the energy is 34.5 meV.

Summary and Future Directions

In this review we have discussed the basic characteristics of magnetic neutron and
X-ray scattering and provided a number of experimental examples of how these
techniques can be employed. Neutron scattering is a rather mature technique which
has the advantage of being a weakly interacting probe that does not affect the
properties of the sample. The source of neutrons has traditionally been steady-state
reactor-based facilities, but this has now been complemented by the newer, pulsed
spallation neutron source facilities which can offer higher peak flux than steady-
state reactors. Both types of sources have many different types of spectrometers that
enable magnetic investigations over many orders-of-magnitude in both spatial and
1292 J. W. Lynn and B. Keimer

time domains. In addition to new sources and new types of sources, many of the
advancements in neutron techniques over the years have come from developments
in how to tailor and manipulate neutrons, vast arrays of detectors, and the software
to analyze and visualize the data, and this progress continues unabated. New sources
and new instrumentation currently are being planned and developed, with the
anticipation that measurement capabilities will be greatly increased together with
an increased quantity and scale of data acquired.
Resonant X-ray scattering is a much newer technique, with high brightness that
allows measurements of small bulk samples, thin films, and multilayers. It also has
the advantage of being element-specific as the resonance is tuned to an absorption
edge. Tremendous progress in measurement capabilities has been realized in the
last few years, both with magnetic diffraction and magnetic inelastic scattering.
In contrast to neutron scattering which is on a solid theoretical foundation,
the theoretical understanding and interpretation of magnetic X-ray scattering is
undergoing considerable development, which should lead to improved interpretation
of experimental data and exciting new capabilities.
The future of both techniques is brilliant, pun intended.

Acknowledgments We thank our many colleagues who have collaborated with us on our own
projects, where we have taken a number of the examples for our own convenience and familiarity
with the systems.

References
1. Lovesey, S.W.: Theory of Neutron Scattering from Condensed Matter. Clarendon Press, Oxford
(1984)
2. Lynn, J.W.: Magnetic neutron scattering. J. Appl. Phys. 75, 6806 (1994)
3. Blume, M.: Magnetic scattering of X-rays. J. Appl. Phys. 57, 3615 (1985)
4. Balcar, E., Lovesey, S.: Theory of Magnetic Neutron and Photon Scattering. Clarendon Press,
Oxford (1989)
5. Hill, J., McMorrow, D.: Resonant exchange scattering: polarization dependence and correlation
functions. Acta Crystallogr. A. 52, 236–244 (1996)
6. Matsumura, T., Nakao, H., Murakami, Y.: Resonant x-ray scattering experiments on the
ordering of electronic degrees of freedom. J. Phys. Soc. Jpn. 82, 021007 (2013)
7. Ament, L., van Veenendaal, M., Devereaux, T.P., Hill, J.P., van den Brink, J.: Resonant inelastic
x-ray scattering studies of elementary excitations. Rev. Mod. Phys. 83, 705 (2011)
8. Hatton, P.D., Johnson, R.D., Bland, S.R., Mazzoli, C., Beale, T.A.W., Du, C.-H., Wilkins, S.B.:
Magnetic structure determination using polarised resonant X-ray scattering. J. Magn. Magn.
Mater. 321, 810 (2009)
9. Moon, R.M., Riste, T., Koehler, W.C.: Polarization Analysis of Thermal-Neutron Scattering.
Phys. Rev. 181, 920–931 (1969)
10. Bacon, G.E.: Neutron Diffraction, 3rd edn. Oxford University Press, Oxford (1975)
11. Blume, M.: Orbital contribution to the magnetic form factor of Ni++ . Phys. Rev. 124, 96 (1961)
12. Trammell, G.T.: Magnetic scattering of neutrons from rare earth ions. Phys. Rev. 92, 1387
(1953)
13. Williams, G.W.: Polarized Neutrons. Clarendon Press, Oxford (1988)
14. Poole, A., Lelievre-Berna, E., Wills, A.S.: General refinement strategy for magnetic structures
using spherical neutron polarimetry and representation analysis. Physica B. 404, 2535 (2009)
25 Magnetic Scattering 1293

15. Zhang, H., Lynn, J.W.: Analytic calculation of polarized neutron reflectivity from supercon-
ductors. Phys. Rev. B. 48, 15893 (1993)
16. Daillant, J., Gibaud, A. (eds.): X-ray and Neutron Reflectivity: Principles and Applications.
Springer, Berlin (2009)
17. Toperverg, B.P., Zabel, H.: Neutron scattering in nanomagnetism. In: Fernandez-Alonso, F.,
Price, D.L. (eds.) Experimental Methods in the Physical Sciences: Neutron Scattering –
Magnetic and Quantum Phenomena, vol. 48, p. 339. Elsevier, London (2015)
18. Haverkort, M.W., Hollmann, N., Krug, I.P., Tanaka, A.: Symmetry analysis of magneto-optical
effects: the case of x-ray diffraction and absorption at the transition metal L2,3 edge. Phys. Rev.
B. 82, 094403 (2010)
19. Leininger, P., Rahlenbeck, M., Raichle, M., Rohnenbuck, B., Maljuk, A., Lin, C.T., Keimer, B.,
Weschke, E., Schierle, E., Seki, S., Tokura, Y., Freeland, J.W.: Electronic structure, magnetic,
and dielectric properties of the edge-sharing copper oxide chain compound NaCu2 O2 . Phys.
Rev. B. 81, 085111 (2010)
20. McWhan, D.B., Isaacs, E.D., Carra, P., Shapiro, S.M., Thole, B.T., Hoshino, S.: Resonant
magnetic s-ray scattering from mixed-valence TmSe. Phys. Rev. B. 47, 8630 (1993)
21. Fleury, P.A., Loudon, R.: Scattering of light by one- and two-magnon excitations. Phys. Rev.
166, 514–530 (1966)
22. Braicovich, L., Ament, L.J.P., Bisogni, V., Forte, F., Aruta, C., Balestrino, G., Brookes, N.B.,
De Luca, G.M., Medaglia, P.G., Granozio, F.M., Radovic, M., Salluzzo, M., van den Brink,
J., Ghiringhelli, G.: Dispersion of magnetic excitations in the cuprate La2 CuO4 and CaCuO2
compounds measured using resonant X-ray scattering. Phys. Rev. Lett. 102, 167401 (2009)
23. Braicovich, L., van den Brink, J., Bisogni, V., Moretti Sala, M., Ament, L.J.P., Brookes, N.B.,
DeLuca, G.M., Salluzzo, M., Schmitt, T., Strocov, V.N., Ghiringhelli, G.: Magnetic excitations
and phase separation in the underdoped La2-x Srx CuO4 superconductor measured by resonant
inelastic x-ray scattering. Phys. Rev. Lett. 104, 077002 (2010)
24. Tacon, M.L., Ghiringhelli, G., Chaloupka, J., Moretti Sala, M., Hinkov, V., Haverkort, M.W.,
Minola, M., Bakr, M., Zhou, K.J., Blanco-Canosa, S., Monney, C., Song, Y.T., Sun, G.L., Lin,
C.T., DeLuca, G., Salluzzo, M., Khaliullin, G., Braicovich, L., Schmitt, T., Keimer, B.: Intense
paramagnon excitations in a large family of high-temperature superconductors. Nat. Phys. 7,
725–730 (2011)
25. Zhou, K.-J., Huang, Y.-B., Monney, C., Dai, X., Strocov, V.N., Wang, N.-L., Chen, Z.-G.,
Zhang, C., Dai, P., Patthey, L., van den Brink, J., Ding, H., Schmitt, T.: Persistent high-energy
spin excitations in iron-pnictide superconductors. Nat. Commun. 4, 1470 (2013)
26. Bisogni, V., Simonelli, L., Ament, L., Forte, F., Moretti Sala, M., Minola, M., Huotari, S., van
den Brink, J., Ghiringhelli, G., Brookes, N., Braicovich, L.: Bimagnon studies in cuprates with
resonant inelastic x-ray scattering at the O K edge. I. Assessment on La2 CuO4 and comparison
with the excitation at Cu L3 and Cu K edges. Phys. Rev. B. 85, 214527 (2012)
27. Kim, J., Daghofer, M., Said, A.H., Gog, T., van den Brink, J., Khaliullin, G., Kim, B.J.:
Excitonic quasiparticles in a spin-orbit Mott insulator. Nat. Commun. 5, 4453 (2014)
28. Suzuki, H., Gretarsson, H., Ishikawa, H., Ueda, K., Yang, Z., Liu, H., Kim, H., Kukusta, D.,
Yaresko, A., Minola, M., Sears, J., Francoual, S., Wille, H.-C., Nuss, J., Takagi, H., Kim, B.,
Khaliullin, G., Yavas, H., Keimer, B.: Spin waves and spin-state transitions in a ruthenate high-
temperature antiferromagnet. Nat. Mater. 18, 563 (2019)
29. Minola, M., Dellea, G., Gretarsson, H., Peng, Y.Y., Lu, Y., Porras, J., Loew, T., Yakhou, F.,
Brookes, N.B., Huang, Y.B., Pelliciari, J., Schmitt, T., Ghiringhelli, G., Keimer, B., Braicovich,
L., Le Tacon, M.: Collective nature of spin excitations in superconducting cuprates probed by
resonant inelastic X-ray scattering. Phys. Rev. Lett. 114, 217003 (2015)
30. Li, X., Ma, X., Su, D., Liu, L., Chisnell, R., Ong, S.P., Chen, H., Toumar, A., Idrobo, J.-C., Lei,
Y., Bai, J., Wang, F., Lynn, J.W., Lee, Y.S., Ceder, G.: Direct visualization of the Jahn-Teller
effect coupled to Na ordering in Na5/8 MnO2 . Nat. Mater. 13, 586 (2014)
31. Huang, Q., Karen, P., Karen, V.L., Kjekshus, A., Migdall, A.D., Rosov, N., Lynn, J.W., Santoro,
A.: Neutron powder diffraction study of the nuclear and magnetic structures of YBa2 Fe3 O8 at
room temperature. Phys. Rev. B. 45, 9611 (1992)
1294 J. W. Lynn and B. Keimer

32. Li, W.-H., Wang, C.-W., Hsu, D., Lee, C.-H., Wu, C.-M., Chou, C.-C., Yang, H.-D., Zhao,
Y., Chang, S., Lynn, J.W., Berger, H.: Interplay between the magnetic and electric degree-of-
freedoms in multiferroics Co3 TeO6 . Phys. Rev. B. 85, 094431 (2012)
33. Lee, C.-H., Wang, C.-W., Zhao, Y., Li, W.-H.J., Harris, A., Rule, K., Yang, H.-D., Berger, H.:
Complex magnetic incommensurability and electronic charge transfer through the ferroelectric
transition in multiferroic Co3 TeO6 . Sci. Rep. 7, 6437 (2017)
34. Wang, C.-M., Lee, C.-H., Li, C.-Y., Wu, C.-M., Li, W.-H., Chou, C.-C., Yang, H.-D., Lynn,
J.W., Huang, Q., Harris, A.B., Berger, H.: Complex magnetic couplings in Co3 TeO6 . Phys.
Rev. B. 88, 184427 (2013)
35. Choi, S.-M., Lynn, J.W., Lopez, D., Gammel, P.L., Canfield, P.C., Bud’ko, S.L.: Direct
observation of spontaneous weak-ferromagnetism in the superconductor ErNi2 B2 C. Phys. Rev.
Lett. 87, 107001 (2001)
36. Thomlinson, W., Lynn, J.W., Shirane, G., Moncton, D.E.: Neutron Scattering Studies of
Magnetic Ordering in Ternary Superconductors, Vol. 34, Chap. 8 ed., M. B. M. a. O. Fischer,
Ed. Springer, New York (1983)
37. Choi, S.-M., Lynn, J.W., Lopez, D., Gammel, P.L., Varma, C.M., Canfield, P.C., Bud’ko, S.L.:
Ferromagnetism and Spontaneous Vortex Formation in Superconducting ErNi2 B2 C. NCNR
annual report 2001. http://www.ncnr.nist.gov/AnnualReport/FY2001_pdf/ (2001)
38. Kawano-Furukawa, H., Ishida, Y., Yano, F., Nagatomo, R., Noda, A., Nagara, T., Ohira-
Kawamura, S., Kobayashi, C., Yoshizawa, K.L., Winn, B.L., Furakawa, N., Takeya, H.:
Creation of vortices by ferromagnetic order in ErNi2 B2 C. Physica C. 470, 5716 (2010)
39. Gardner, J.S., Gingras, M.J.P., Greedan, J.E.: Magnetic pyrochlore oxides. Rev. Mod. Phys. 82,
53 (2010)
40. Bramwell, S.T., Harris, M.J., den Hertog, B.C., Gingras, M.J.P., Gardner, J.S., McMorrow,
D.F., Wildes, A.R., Cornelius, A.L., Champion, D.M., Melko, R.G., Fennell, T.: Spin correla-
tions in Ho2 Ti2 O7 : a dipolar spin ice system. Phys. Rev. Lett. 87, 047205 (2001)
41. Kadowaki, H., Doi, N., Aoki, Y., Tabata, Y., Sato, T.J., Lynn, J.W., Matsuhira, K., Hiroi, Z.:
Observation of magnetic monopoles in spin ice. J. Phys. Soc. Jpn. 78, 103706 (2009)
42. Costelnovo, C., Moessner, R., Sondhi, S.L.: Magnetic monopoles in spin ice. Nature. 451, 42
(2008)
43. Morris, D.J.P., Tennant, D.A., Grigera, S.A., Klemke, B., Castelnovo, C., Moessner, R.,
Cztemasty, C., Meissner, M., Rule, K.C., Hoffmann, J.U., Kiefer, K., Gerischer, S., Slobinsky,
D., Perry, R.S.: Dirac strings and magnetic monopoles in the spin ice Dy2 Ti2 O7 . Science. 326,
411 (2009)
44. Fennell, T., Deen, P.P., Wildes, A.R., Schmalzl, K., Prabhakaran, D., Boothroyd, A.T., Aldus,
R.J., McMorrow, D.F., Bramwell, S.T.: Magnetic coulomb phase in the spin ice Ho2 Ti2 O7 .
Science. 326, 415 (2009)
45. Santos, T.S., Kirby, B.J., Kumar, S., May, S.J., Borchers, J.A., Maranville, B.B., Zarestky, J.,
Te Velthuis, S.G.E., van den Brink, J., Bhattacharya, A.: Delta doping of ferromagnetism in
antiferromagnetic manganite superlattices. Phys. Rev. Lett. 107, 167202 (2011)
46. Kim, B.J., Ohsumi, H., Komesu, T., Sakai, S., Morita, T., Takagi, H., Arima, T.: Phase sensitive
observation of a spin-orbital Mott state in Sr2 IrO4 . Science. 323, 1329 (2009)
47. Garcia-Fernandez, M., Wilkins, S.B., Lu, M., Qing’an, L., Gray, K.E., Zheng, H., Mitchell,
J.F., Khomskii, D.: Antiferromagnetic domain structure in a bilayer manganite. Phys. Rev. B.
88, 075134 (2013)
48. Frano, A., Schierle, E., Haverkort, M.W., Lu, Y., Wu, M., Blanco-Canosa, S., Nwankwo, U.,
Boris, A.V., Wochner, P., Cristiani, G., Habermeier, H.U., Logvenov, G., Hinkov, V., Benckiser,
E., Weschke, E., Keimer, B.: Orbital control on noncollinear magnetic order in nickel oxide
heterostructures. Phys. Rev. Lett. 111, 106804 (2013)
49. Macke, S., Hamann-Borrero, J.E., Green, R.J., Keimer, B., Sawatzky, G.A., Haverkort, M.W.:
Dynamical effects in resonant X-ray diffraction. Phys. Rev. Lett. 117, 115501 (2016)
25 Magnetic Scattering 1295

50. Forst, M., Caviglia, A., Scherwitzl, R., Mankowsky, R., Zubko, P., Khanna, V., Bromberger, H.,
Wilkins, S.B., Chuang, Y.D., Lee, W.S., Schlotter, W.F., Tumer, J.J., Dakovski, G.L., Minitti,
M.P., Robinson, J., Clark, S.R., Jaksch, D., Triscone, J.M., Hill, J.P., Dhesi, S.S., Cavalleri,
A.: Spatially resolved ultrafast dynamics initiated at a complex oxide interface. Nat. Mater. 14,
883 (2015)
51. Lynn, J.W., Erwin, R.W., Borchers, J.A., Huang, Q., Santoro, A., Peng, J., Li, Z.: Unconven-
tional ferromagnetic transition in La1-x Cax MnO3 . Phys. Rev. Lett. 76, 4046 (1996)
52. Lynn, J.W., Rosov, N., Fish, G.: Polarization analysis of the magnetic excitations in invar and
non-invar amorphous ferromagnets. J. Appl. Phys., 73, 5369–5371 (1993)
53. Vasiliu-Doloc, L., Lynn, J.W., Moudden, A.H., de Leon-Guevara, A.M., Revcolevschi, A.:
Structure and spin dynamics of La0.85 Sr0.15 MnO3 . Phys. Rev. B. 58, 14913 (1998)
54. Ye, F., Dai, P., Fernandez-Baca, J.A., Sha, H., Lynn, J.W., Kawano-Furukawa, H., Tomioka, Y.,
Tokura, Y., Zhang, J.: Evolution of spin-wave excitations in ferromagnetic metallic manganites.
Phys. Rev. Lett. 96, 047204 (2006)
55. Bayrakci, S.P., Tennant, D.A., Leinninger, P., Keller, T., Gibson, C.R., Wilson, S.D., Habicht,
K., Birgeneau, R.J., Keimer, B.: Lifetimes of antiferromagnetic magnons in two and three
dimensions: experiment, theory, and numerics. Phys. Rev. Lett. 111, 017204 (2006)
56. Fujita, M., Hiraka, H., Matsuda, M., Matsuura, M., Tranquada, J.M., Wakimoto, S., Xu, G.,
Yamada, K.: Progress in neutron scattering studies of spin excitations in high-Tc cuprates. J.
Phys. Soc. Jpn. 81, 011007 (2012)
57. Lumsden, M.D., Christianson, A.D.: Magnetism in Fe-based superconductors. J. Phys. Con-
dens. Matter. 22, 203203 (2010)
58. Harriger, L.W., Luo, H.Q., Liu, M.S., Frost, C., Hu, J.P., Norman, M.R., Dai, P.: Nematic spin
fluid in the tetragonal phase of BaFe2 As2 . Phys. Rev. B. 84, 054544 (2011)
59. Lake, B., Tennant, D.A., Caux, J.-S., Barthel, T., Schollwock, U., Nagler, S.E., Frost, C.D.:
Multispinon continua at zero and finite temperature in a near-ideal Heisenberg chain. Phys.
Rev. Lett. 111, 137205 (2013)
60. Mourigal, M., Enderle, M., Klöpperpieper, A., Caux, J.-S., Stunault, A., Ronnow, H.M.:
Fractional spinon excitations in the quantum Heisenberg antiferromagnetic chain. Nat. Phys.
9, 435 (2013)
61. Gretarsson, H., Saha, S.R., Drye, T., Paglione, J., Kim, J., Casa, D., Gog, T., Wu, W., Julian,
S.R., Kim, Y.-J.: Spin state transition in the Fe pnictides. Phys. Rev. Lett. 110, 047003 (2013)
62. Kreyssig, A., Green, M., Lee, Y., Samolyuk, G., Zajdel, P., Lynn, J., Bud’ko, S., Torikachvili,
M., Ni, N., Nandi, S., Leao, J., Poulton, S., Argyriou, D., Harmon, B., McQueeney, R.,
Canfield, P., Goldman, A.: Pressure-induced volume-collapsed tetragonal phase of CaFe2 As2
as seen via neutron scattering. Phys. Rev. B. 78, 184517 (2008)
63. Saha, S., Butch, N., Drye, T., Magill, J., Ziemak, S., Kirshenbaum, K., Lynn, J. W., Paglione,
J.: Structural collapse and superconductivity in rare-earth-doped CaFe2 As2 . Phys. Rev. B. 85,
024525 (2012)

Jeffrey Lynn received his Ph.D. in 1974 from Georgia Tech while
conducting his thesis research at Oak Ridge National Laboratory.
He was a postdoctoral Fellow at Brookhaven National Laboratory,
before becoming Professor of Physics, University of Maryland
and consultant at the National Bureau of Standards (now NIST).
Currently he is a NIST Fellow and Adjunct Professor in the
Quantum Materials Center at the University of Maryland.
1296 J. W. Lynn and B. Keimer

Bernhard Keimer received his Ph.D. in 1991 from the Mas-


sachusetts Institute of Technology. Upon graduating he joined
the faculty of the Department of Physics of Princeton University,
becoming Full Professor in 1997. In 1998 he became Director of
the Max Planck Institute for Solid State Research in Stuttgart and
is a Member of the Scientific Council, Max Planck Society.
Electron Paramagnetic and Ferromagnetic
Resonance 26
David Menard and Robert Barklie

Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1298
Magnetic Resonance of Electrons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1299
Microscopic Equation of Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1299
Macroscopic Equation of Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1301
On the Difference Between EPR and FMR . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1302
Electron Paramagnetic Resonance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1303
Bloch Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1303
Effective Spin Hamiltonian . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1304
Anisotropic Zeeman Interaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1306
Fine Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1308
Hyperfine Interaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1310
Ferromagnetic Resonance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1312
Effective Magnetic Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1312
Shape Anisotropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1314
Magnetocrystalline Anisotropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1316
Exchange Energy and Spin Wave Resonance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1317
FMR in Metals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1319
Experimental Observation of Electron Magnetic Resonance . . . . . . . . . . . . . . . . . . . . . . . . . 1321
Frequency Domain Cavity EPR/FMR . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1321
Survey of Other Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1324
Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1329
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1329
Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1331

D. Menard ()
Department of Engineering Physics, Polytechnique Montreal, Montréal, QC, Canada
e-mail: david.menard@polymtl.ca
R. Barklie
School of Physics, Trinity College, Dublin, Ireland

© Springer Nature Switzerland AG 2021 1297


J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_25
1298 D. Menard and R. Barklie

Abstract

This chapter presents the principles and practice of electron paramagnetic reso-
nance (EPR) and of ferromagnetic resonance (FMR). In spite of their common
physical origin and experimental methods, EPR and FMR are rarely presented
together. The combined treatment emphasizes their similarities and differences.
EPR is treated quantum mechanically in terms of transitions between energy
levels, exploiting the concepts of effective spin Hamiltonian and a rotating frame
of reference. FMR uses a macroscopic method to describe quasi-rigid macro-
scopic magnetization with precessional motion around an effective magnetic
field. Both approaches involve phenomenological parameters in a mathematical
form consistent with the symmetry of the interactions. The parameters, which
can be calculated theoretically in principle, are often determined experimentally
by fitting the angle-dependent data to the calculated phenomenological response.
The classical methods of observing EPR and FMR are presented, along with a
survey of alternative complementary techniques.

Abbreviations
AFC Automatic frequency control
AWG Arbitrary waveform generator
CPW Coplanar waveguide
CW Continuous wave
EDMR Electrically detected magnetic resonance
ENDOR Electron nuclear double resonance
EPR Electron paramagnetic resonance
FMR Ferromagnetic resonance
GMI Giant magnetoimpedance
HF Hyperfine
LLG Landau-Lifshitz-Gilbert
LNA Low-noise amplifier
NMR Nuclear magnetic resonance
ODMR Optically detected magnetic resonance
PIMM Pulsed inductive microwave magnetometry
PSD Phase-sensitive detection
SHE Spin Hall effect
SHF Super hyperfine
TR-MOKE Time-resolved magneto-optical Kerr effect
TWTA Traveling-wave tube amplifier
VNA Vector network analyzer

Introduction

The observation of nuclear magnetic resonance (NMR) [1, 2], electron paramagnetic
resonance (EPR) [3, 4], and ferromagnetic resonance (FMR) [5] was reported in the
mid-1940s, following declassification of radar and microwave research that took
place during World War II. By 1950, most of the underlying physical principles
of the three resonance phenomena were well understood, except for the detailed
26 Electron Paramagnetic and Ferromagnetic Resonance 1299

relaxation mechanisms leading to resonance line broadening. From there on, the
three domains have evolved into distinct communities, gradually developing their
own languages, formalisms, and powerful characterization techniques associated
with various areas of physics, chemistry, and materials sciences.
This chapter gives a brief, simplified account of the principles and practice
of EPR and FMR spectroscopy. Comprehensive accounts are provided in several
excellent EPR [6–10] and FMR [11–13] textbooks. Our aim is to explain the
occurrence of electron magnetic resonance in paramagnetic and ordered materials,
highlighting the common physical origin and the distinct formalism of the two
methods. We do not cover the relaxation mechanisms responsible for the resonance
linewidth, for reasons of conciseness. References are kept at a minimum, mostly
focusing on textbooks and review chapters.

Magnetic Resonance of Electrons

Microscopic Equation of Motion

Magnetic resonance results from the proportionality between an elementary mag-


netic dipole moment and the associated angular momentum. When such a combi-
nation of magnetic and kinetic moments is placed in a constant uniform magnetic
field B0 , it experiences a torque, which results in a precession about B0 at a rate
proportional to B0.
In quantum mechanics,
  the expectation value of the angular momentum is

quantized, that is, Jˆ =  J (J + 1), where J is an integral or half-integral
quantum number. We follow here the widespread practice of using dimensionless
angular momentum operator (DM apologizes). The projection of the moment μz
on an arbitrary quantification axis, conventionally chosen to be the direction of the
uniform magnetic field B0 parallel
  to the z axis, is related to the quantized angular
momentum along that axis, Jˆz = MJ by

μz = −gμB MJ = γ MJ , (1)

where MJ is a quantum number taking integral or half-integral values, μB = |e|/2me


is the Bohr magneton, γ = − g|e|/2me is the gyromagnetic ratio, g is the
dimensionless spectroscopic splitting factor,  is Planck’s constant, and e and me
are the electron charge and mass, respectively. The magnetic part of the Hamiltonian
Ĥ = −μ̂z B0 = −γ Jˆz B0 yields the equation of motion for the expectation value
of the moment,
 
d μ̂ i   i  
= Ĥ , μ̂ = − γ B0 Jˆz , γ Jˆ , (2)
dt  
1300 D. Menard and R. Barklie

which can be rewritten using the commutation relations for angular momentum [8]:
 
d μ̂  
= − |γ | μ̂ × B 0 . (3)
dt

The absolute value for the gyromagnetic ratio has been chosen to remove any
ambiguity on the sign of γ , as both signs are used in the literature. Solutions of (3)
describe a classical precession of the moment μ around B0 at an angular frequency:

ω0 = |γ | B0 , (4)

also called the Larmor frequency. Alternatively, (3) may be obtained by combining
the classical equation of motion (with J in units of ) dJ/dt = μ × B0 with the
relation μ = γ J leading to the same equation of motion, provided one replaces the
expectation value by a proper statistical average.
It is common practice in EPR and NMR to employ an alternative coordinate sys-
tem, rotating
  at an angular frequency  
 around the z axis, such that (3) transforms
into d  μ̂ /dt = (|γ | B 0 − ) × μ̂ [8]. As a result, in the rotating frame, the
magnetic moment precesses around the effective field B e = (B0 − / |γ |) ẑ and
remains fixed for  = ω0 . By including a microwave field B1 alongside B0 in the
Hamiltonian, where B1 is circularly polarized at right angles to B0 and oscillating at
a frequency of ω0 , so that B1 is fixed in the frame of reference rotating at  = ω0 ,
the solution of (2) or (3) yields a periodic oscillation of μ around B1 , called Rabi
oscillations, at a frequency of ω1 = |γ |B1 in the rotating frame of reference.
Quantum mechanics shows that the effect of the B0 field is to split the degenerate
ground state of the system into 2MJ + 1 equally spaced energy levels, separated by

E = ω0 =  |γ | B0 = gμB B0 . (5)

The resonance is often pictured as a transition between adjacent energy levels,


as the system is submitted to the microwave field B1 . For example, if the moment
arises only from the electron’s spin angular momentum, so that MJ = Ms = ± 1/2
and g is approximately 2, the ground state splits in two levels which are separated
by ΔE = 2μB B0 , and the resonance frequency is given by (4), where |γ |/2π is
approximately 28 GHz/T (Fig. 1):

Fig. 1 Zeeman splitting of a


single spin
26 Electron Paramagnetic and Ferromagnetic Resonance 1301

In practice, it is technically easier to generate a linearly polarized B1 field,


which is equivalent to two circularly polarized fields rotating in opposite direc-
tions, only one of which causes the transitions. If B1 is parallel to x, then the
transition rate between the two states |MJ , |MJ ± 1 will be proportional to

2
 
MJ | Jˆ+ + Jˆ− |MJ ± 1 which has the same value in both directions; fur-
thermore, this rate is zero unless the frequency of B1 is such that the resonance
condition (5) is verified. If the individual moments were isolated from each other,
the probability per unit time of the transition between up and down states being the
same in both directions implies that the effect of B1 alone at resonance would be to
make their populations become equal. Once they are equal, there will be no further
power absorption by the sample and no EPR signal, but this does not happen in
practice because of the effects of relaxation.

Macroscopic Equation of Motion

In an EPR or FMR experiment, the sample frequently contains a large number of


magnetic moments. These moments are usually not isolated from each other or from
other degrees of freedom such as lattice vibrations. As a result of the interactions
with their environment, the moments will experience additional torques, which may
be treated phenomenologically as effective magnetic fields. Also, the excess energy
in the spin system will be transferred to these other degrees of freedom at various
rates. In these circumstances, it is more appropriate to discuss the bulk properties
of the spin system rather than the individual magnetic moments and therefore to
modify (3) to account for the existence of effective fields and relaxation processes.
If there are N elementary magnetic moments in a volume V, and the average
macroscopic magnetization is defined as

1  
N
M= μi , (6)
V
i=1

then the equation of motion can be rewritten as

dM
= − |γ | M × B eff − R, (7)
dt

where Beff is an effective field and R is a phenomenological relaxation term to


be discussed below. It is worth mentioning that Beff , as it is used here, does not
represent the field in the rotating frame as often encountered in EPR literature. It
actually accounts for all interactions that result in a torque on the magnetization,
including those with B0 + B1 . In (6) M should be understood as resulting from a
proper ensemble average as prescribed by statistical physics.
The relaxation term R can take different forms. Here we only consider the two
most common, which are referred to as the Bloch relaxation terms,
1302 D. Menard and R. Barklie


(Mz − M0 ) ẑ Mx x̂ + My ŷ
RB = + , (8)
T1 T2

usually associated with NMR and EPR studies, and the Gilbert relaxation term,

 
α dM
RG = M× , (9)
Ms dt

mostly used for FMR. We discuss them further in Sects. “Electron Paramagnetic
Resonance” and “Ferromagnetic Resonance”.

On the Difference Between EPR and FMR

EPR and FMR are both related to electron resonance: in paramagnetic substances
for EPR and in materials with magnetic order for FMR. Both are based on the same
physics, as described by (7), and can be observed with the same experimental setup.
What differentiates EPR and FMR is the nature and magnitude of the interactions
experienced by the elementary moments, as phenomenologically included in the
effective fields in (7). Historically, their practitioners adopted their own theoretical
formalisms to treat these interactions: for EPR microscopically in terms of an
effective spin Hamiltonian (Sect. “Electron Paramagnetic Resonance”) and for
FMR by exploiting a macroscopic micromagnetic free energy (Sect. “Ferromagnetic
Resonance”).
The usual aim of an EPR experiment is to obtain information about the nature
and population of paramagnetic centres by inducing magnetic dipole transitions
between the energy levels. These are associated with the local environment of the
magnetic moment, including the presence of nearby nuclear spins or interaction
with the crystal field. A resonant centre may be an impurity atom, a free radical, or
a defect such as a vacancy; these are often referred to as point defects since they have
zero dimension. The local environment provides small perturbations of the energy
levels with respect to that of the isolated moment. The information about the centres
is extracted from the number and positions of the resonance lines, as well as their
shapes and linewidths.
In FMR studies, the specimen may be viewed as an ensemble of magnetic
moments locked together by the strong, short-range, exchange coupling. As such,
they tend to precess together as a large macrospin in an effective mean field.
The interactions of the moments with their environment are no longer small
perturbations, as they often result in effective fields far greater than the applied
fields. Exchange and dipolar interactions also give rise to a rich spectrum of spin
wave excitations. The resonance positions and linewidths enable us to probe spin
wave excitations and magnetization dynamics, and they may serve as chemical and
structural characterization parameters.
26 Electron Paramagnetic and Ferromagnetic Resonance 1303

Electron Paramagnetic Resonance

Bloch Equations

The combination of (7) and (8), which was used by Bloch [14] to discuss early
NMR experiments, will be referred to as the Bloch equations. In (8), M0 represents
the thermal equilibrium value of the magnetization. The spin-lattice relaxation time,
T1 , is therefore associated with the relaxation of the longitudinal component of the
magnetization toward thermal equilibrium, whereas the spin-spin relaxation time,
T2 , describes the loss of coherence of the transverse components. As noted by Pake
and Estle [8], these equations give the simplest description for magnetic resonance
when relaxation is included, although, as they point out, many paramagnetic
samples will not obey these equations. They are useful because their solutions
provide a guide to understanding what often happens and they introduce the
important concepts of dispersion and absorption.
As a quick primer to EPR, let us consider the time dependence of the magneti-
zation in a continuous wave (CW) experiment as described by the Bloch equations.
Solving (7) combined with (8), when the effective field reduces to the applied field,
B(t) = B0 ẑ + B1 e−iωt x̂, yields [8]

μ0 M⊥ = χ  − iχ  B1 , (10)

where

(ω0 − ω) ωM T22
χ = , (11)
1 + (ω0 − ω)2 T22 + ω12 T1 T2

ωM T 2
χ  = , (12)
1 + (ω0 − ω)2 T22 + ω12 T1 T2

for the amplitude of the rotating transverse magnetization and


 
1 + (ω0 − ω)2 T22
Mz = M0 , (13)
1 + (ω0 − ω)2 T22 + ω12 T1 T2

for the longitudinal component. In (10–13), μ0 is the permeability of free space,


ωM = |γ |μ0 M0 , and ω0 and ω1 are associated with B0 and B1 as defined in Sect.
“Microscopic Equation of Motion”. It follows from (11) and (12) that the real and
imaginary parts of the complex susceptibility, χ and χ , exhibit a typical dispersion
and absorption behavior, respectively, as shown in Fig. 2.
Very often, the quantity accessible from a magnetic resonance experiment is the
average power P per unit volume absorbed by the spin system from the microwave
field [6–9]:
1304 D. Menard and R. Barklie

Fig. 2 Dispersive and


dissipative components of the
magnetic susceptibility

Applied field

ωχ  B12
P = . (14)
μ0

Equation (12) shows that χ , and hence P, has a Lorentzian shape with a
maximum at ω = ω0 , given by

 ωM T2
χmax = , (15)
1 + ω12 T1 T2

and a full linewidth at half amplitude of



2
B1/2 = 1 + ω12 T1 T2 . (16)
γ T2

From Eqs. (12) and (14), it follows that


 
ωωM T2 B12
P = . (17)
1 + (ω0 − ω)2 T22 + ω12 T1 T2 μ0

This shows that P initially increases as B1 increases but that saturation occurs
when ω12 T1 T2  1. In the absence of saturation, for small values of B1 , such that ω1
is small compared to relaxation rates , that is, ω12 T1 T2  1, we see that Mz M0 ,
the linewidth simplifies to ΔB1/2 = 2/|γ |T2 , and the maximum of value of χ at ω0
 = ω T . Another important result, as noted elsewhere [9], is that,
is given by χmax M 2
in the absence of saturation, the area under the χ  curve is independent of T1 and T2
and is proportional to M0 and, hence, to the number of spins per unit volume, N/V.

Effective Spin Hamiltonian

The EPR spectrum of a single isotropic line with the free electron g-value of
2.002319 is rarely observed. Instead EPR spectra frequently display one or more
lines whose field positions depend on the direction of the applied magnetic field
26 Electron Paramagnetic and Ferromagnetic Resonance 1305

relative to the sample. The aim of this section is to describe several interactions
which give rise to these effects, to show how the relevant parameters may be
extracted from the spectra, and to explain what these reveal about the nature of
the centres.
The field positions of the lines in the spectra come from the resonance condition
ω = ΔE, where ΔE is the energy difference between each pair of levels between
which an electron magnetic dipole transition can occur and ω is the microwave
angular frequency of B1 . To find ΔE, it is necessary to determine the expressions
for the appropriate energy levels. We are primarily interested in energy levels close
to the ground state, since the photon energies used for achieving EPR are usually
only a few cm−1 (on the order of μeV), whereas the ground states are separated
from other excited states by energies much greater than this.
We often refer to this group of states as the ground manifold. As they transform
like spin angular eigenfunctions, we associate an effective spin S̃ to the ground
manifold to represent the behavior of this group of states when a magnetic field is
applied. This is a fictitious spin angular momentum such that the degeneracy of this
group of levels is set equal to (2S̃ + 1). The Hamiltonian operator associated with
this effective spin is called the effective spin Hamiltonian. S̃ may sometimes be the
true spin S, and in what follows we shall continue the usual practice of omitting the
difference between S̃ and S and refer only to the spin Hamiltonian, even though an
orbital component may contribute to it.
Hence, the energy levels appropriate to EPR are the eigenvalues of the eigen-
functions of the relevant spin Hamiltonian. A general and practical form for the spin
Hamiltonian is given by [8]

ĤS = ĤZ + ĤFS + ĤHF + ĤNZ + ĤQ . (18)

The first term is the essential part of the paramagnetic Hamiltonian. It is


associated with the effective electron Zeeman splitting,

ĤZ = μB S · g · B 0 , (19)

where g is a second-order tensor. The second term

ĤFS = S · D · S (20)

is associated with the fine structure interaction, where D is the fine structure tensor,
and the other terms, which contain the nuclear spin operator I,

ĤHF = S · A · I , ĤNZ = −μn B · g n · I , ĤQ = I · Q · I , (21)

characterize the hyperfine structure, where the A tensor is associated with the
magnetic hyperfine interaction; the second term accounts for the Zeeman splitting
of the nuclear spin and the last one for the nuclear electric quadrupole interaction.
1306 D. Menard and R. Barklie

The appropriate form of the spin Hamiltonian (18–21) is specific for each system.
It is possible in principle to calculate its mathematical form, using perturbation
theory or other means, but it is often more convenient and quicker to work with
the phenomenological expression, which covers a majority of the cases encountered
in practice. Simplifications arise by choosing parameters compatible with the
symmetry of the system under study, and the actual values of those parameters
are then deduced from the EPR spectra. In what follows, we discuss how the
characteristics of the EPR spectra are determined by the Zeeman, fine structure, and
hyperfine structure interactions; the description of the nuclear electric quadrupole
interaction and its effect on the EPR spectrum can be found elsewhere [6–9].

Anisotropic Zeeman Interaction

Unlike the electron Zeeman interaction of a free electron treated in Sect. “Bloch
Equations”, for most paramagnetic centres, the interaction depends also on the
direction of the field with respect to certain axes defined by the local symmetry
of the centre. In these cases, the Hamiltonian is given by (19), and the g tensor,
represented by a 3 × 3 matrix, reflects the local symmetry of the centre. In most
practical cases, by choosing suitable axes x, y, z, known as the principal axes, the
tensor becomes diagonal, yielding


ĤZ = μB gxx Sx Bx + gyy Sy By + gzz Sz Bz . (22)

If the direction of the magnetic field is defined by the direction cosines (α x , α y ,


α z ) with respect to the principal axes, then the Zeeman energy becomes

E = gμB B0 MS (23)

where

g= 2 α2 + g2 α2 + g2 α2.
gxx x yy y zz z (24)

Calculation of the transition probabilities for the magnetic dipole transitions


shows that the selection rule is ΔMS = ± 1 and hence the resonance condition
ΔE = ω = gμB B0 given by (5) remains valid, provided g is defined by (24), so
that for EPR at a constant frequency, the anisotropic resonance is at field values
given by


B0 αx , αy , αz =  . (25)
2 α + g2 α + g2 α
μB gxx x yy y zz z
26 Electron Paramagnetic and Ferromagnetic Resonance 1307

In particular, for an axial symmetry axis parallel to z, gzz = g


, gxx = gyy = g⊥ ,

1/2
and g = g
2 cos2 θ + g⊥ 2 sin2 θ , where θ is the angle between B0 and the axis
(Fig. 3). For cubic symmetry gxx = gyy = gzz and g is isotropic.
The anisotropy of the g factor arises from the unpaired electron having a
component of orbital angular momentum since spin-only angular momentum would
cause g to be isotropic with a value close to ge . The existence of a significant
orbital contribution to the magnetic moment is common for isolated atoms (neutral
or ionized). For instance, in situations where the spin-orbit interaction is smaller
in magnitude than the exchange interaction between the spins, the existence of the
orbital contribution can result in a g-value given by the Lande factor:

J (J + 1) + S (S + 1) − L (L + 1)
gJ = 1 + (26)
2J (J + 1)

that differs significantly from ge . For instance , an atom with a single unpaired
electron in the p shell with L = 1, S = 1/2 would have g = 2/3 and g = 4/3 for
the 2 P1/2 and 2 P3/2 states, respectively. However, when such an atom is within a
solid, the orbital contribution to the magnetic moment is often largely removed,
and the moment arises almost entirely from the electron spin. This quenching of the
angular momentum arises because it is energetically favorable for the electrons to be
in states where they avoid any negatively charged ions or ligands as far as possible.
As an example, Spaeth, Niklas, and Bartram [9] considered an atom with one
unpaired outer p electron (an L = 1, S = 1/2 state) at a site of orthorhombic
symmetry, in which the appropriate atomic states are |px , |py , |pz  states with
energies changed by −, 0, + , respectively, as a result of the electrostatic
interaction with the ions. The ground state, relevant for EPR, is |px , and as this
has no orbital angular momentum, the g-value might be expected to be equal to ge .
However, this ignores the admixture of the excited states into the ground state as a

Fig. 3 (a) Zeeman split energy levels as a function of applied magnetic field and (b) angular
dependence of the resonance line position as a function of field angle θ for an S = 1/2 centre with
axial g tensor
1308 D. Menard and R. Barklie

result of the spin-orbit interaction operator λL · S. The admixture is such that there
remains a small orbital contribution to the magnetic moment. Using perturbation
theory, they showed that [9]

λ 2λ
gxx = ge , gyy = ge − , gzz = ge − , (27)
 

where terms in (λ/Δ)2 are neglected (it is assumed that λ  Δ). In general, as
shown elsewhere [7],
G| Li |n n| Li |G
gij = ge δij − 2λ . (28)
En − EG
n =G

The orbitally nondegenerate ground state is |G, MS  where G represents the


spatial wave function, Li and Lj are orbital angular momentum operators appropriate
to the x, y, z directions, and the summation is over all excited states.
The above description shows that for a paramagnetic centre in a solid, in
situations where the ground state is orbitally nondegenerate and also where λ  Δ,
the observation that g = ge + δg, with δg small, but anisotropic, allows one to
deduce the following information: firstly, that the local symmetry of the centre is
given by that of the g tensor; secondly, that the centre is “electron”- or “hole”- like
when δg is negative (λ is positive) or positive, respectively; and, thirdly, that the
magnitude of δg reflects the magnitude of λ/Δ and is therefore likely to be larger
for higher atomic number elements.

Fine Structure

When a paramagnetic centre has S > 1/2, there may be an interaction called the fine
structure (FS) interaction which leads to additional zero-field splitting of the energy
levels (no applied magnetic field). This can lead to extra lines in the EPR spectrum.
Since for S > 1/2 there is more than one unpaired electron, it is to be expected that
magnetic dipole-dipole interactions between them will occur. However, the situation
can be complicated as the electrical crystal field and spin-orbit interaction may play
a role too; a comprehensive treatment is given by Abragam and Bleaney [6]. An
excellent introduction to this topic is given by Spaeth, Niklas, and Bartram [9].
To illustrate the effect of the FS interaction, let us consider the case where the
paramagnetic centre has S > 1/2 and is in a site of less than cubic symmetry. We
assume a spin Hamiltonian of the form

ĤS = μB S · g · B + S · D · S. (29)
26 Electron Paramagnetic and Ferromagnetic Resonance 1309

With respect to the principal axes x, y, z of the tensor D , the FS term becomes
ĤFS = Dxx Sx2 + Dyy Sy2 + Dzz Sz2 . The trace of D is usually set to 0 since this does
not affect the EPR spectrum. This gives
   
1
ĤFS = D Sz2 − S (S + 1) + E Sx2 − Sy2 , (30)
3

where D = 32 Dzz is the axially symmetric part and E = 12 Dxx − Dyy represents
the deviation from axial symmetry.
  If, to further simplify , the symmetry is axial
(axis//z), g is isotropic, and ĤZ  ĤFS , then to first order in perturbation theory,
the energy levels are given by

1   1

E = gμB B0 MS + D 3cos2 θ − 1 MS2 − S (S + 1) , (31)
2 3

where θ is the angle between B0 and z. In this strong field limit, the eigenfunctions
are |MS . Considering S = 1, MS = −1, 0, 1, and the strongly allowed transitions
for ΔMS = ± 1, an EPR spectrum at a constant frequency yields two lines at field
values given by

ω D  
B0 = ± 3cos2 θ − 1 , (32)
gμB 2gμB

as shown in Fig. 4. Measurement of the angular dependence of the spectrum would


give the values of both D and g (Fig. 4b). For S ≥ 1, there would be 2S lines.

Fig. 4 (a) Energy levels for an S = 1 system as a function of applied field at θ = 0, with zero-
field splitting of D. The two transitions at fixed frequency correspond to the two resonance fields.
(b) Angular dependence of the two resonances which coincide at θ c = 54.74◦
1310 D. Menard and R. Barklie

Hyperfine Interaction

The interaction between the magnetic moment of the unpaired electron (or hole)
of a paramagnetic centre and that of its central nucleus is called the magnetic
hyperfine (HF) interaction. When the interaction is with the magnetic moment of
a neighboring nucleus, it is known as the superhyperfine (SHF) interaction. The
nuclear Zeeman interaction must also be included. Referring to (21), without the
quadrupole term, and assuming an isotropic nuclear Zeeman effect, which is the
most frequently encountered, the spin Hamiltonian including HF interaction is given
by

ĤS = μB S · g · B 0 + S · A · I − gn μn I · B 0 . (33)

The nuclear magnetic moment is related to the nuclear spin I by μ = gn μn I,


where μn is the nuclear magneton and A is the hyperfine tensor.
Assuming that only the first two terms in (33) are significant and that the
HF interaction is small compared to the Zeeman interaction, and for simplicity
considering an orthorhombic symmetry in which the g and HF tensors have the
same principal axes x, y, z, in which the direction of the magnetic field is defined by
the direction cosines (α x , α y , α z ) with respect to these axes, the energy levels are
given to first order by [6–9]

E = gμB B0 MS + KM S MI , (34)

where the nuclear quantum numbers are MI = {−I, . . . , I − 1, I}, the angle-
dependent g factor is given by (24), and

2 α 2 + A2 g 2 α 2 + A2 g 2 α 2
A2xx gxx x yy yy y zz zz z
K=  . (35)
2 α2 + g2 α2 + g2 α2
gxx x yy y zz z

Calculation of the EPR transition probabilities shows that strong (allowed)


transitions occur for ΔMS = ± 1, ΔMI = 0 but that weak (forbidden) ones also can
be seen for ΔMS = ± 1, ΔMI = ± 1, under certain conditions. Thus, the allowed
EPR lines will occur at field values given by

ω K
B0 = − MI . (36)
gμB gμB

Therefore, the presence of an HF interaction is revealed by the fact that there are
2I + 1 lines with equal spacing to first order in K/gμB B0 . Inclusion of second-
order terms makes the spacing depend slightly on MI . The generalization to axial

1/2
symmetry is straightforward by using as previously g = g
2 cos2 θ + g⊥ 2 sin2 θ
26 Electron Paramagnetic and Ferromagnetic Resonance 1311


1/2
and K = g −1 A2
g
2 cos2 θ + A2⊥ g⊥ 2 sin2 θ . In the isotropic case, g and K = A
become scalars.
A further characteristic of such spectra, for interaction with a single nucleus, is
that all the lines have very nearly the same intensity. This occurs because, under
normal conditions of applied field and temperature, the populations of the 2I + 1
different nuclear spin states are almost equal. However, if there are N equivalent
nuclei, each of spin I, there will be 2NI + 1 lines with intensity ratios reflecting the
probability of occurrence of each value of magnetic field produced by the nuclear
moments and experienced by the unpaired electron. For example, for N = 2 and
I = 1/2, there will be three lines reflecting the three field values associated with the
nuclear spin states |↑↑, |↑↓, |↓↑, |↓↓ with relative intensities 1:2:1; the field
produced by both states |↑↓ and |↓↑ is zero.
Let us treat the nuclear Zeeman interaction by adding the third term in (33),
which was neglected before. Assuming orthorhombic symmetry, we can write to
first order, if the nuclear Zeeman effect is small compared to the HF term [8]
 
Axx gxx αx2 + Ayy gyy αy2 + Azz gzz αz2
E = gμB B0 MS + KM S MI − gn μn B0 MI ,
Kg
(37)

where K is given by (35) and g by (24). Likewise, if the Zeeman effect is larger than
the HF splitting, we obtain
 
Axx gxx αx2 + Ayy gyy αy2 + Azz gzz αz2
E = gμB B0 MS − gn μn B0 MI + MS MI .
g
(38)

The generalization to axial and isotropic cases is straightforward.


A simple example of an isotropic hyperfine interaction is that of atomic hydro-
gen, Ho , which can exist not only in the gas phase but also in solids. Many examples
of the latter are listed by Pake and Estle [8]. Other examples involve the EPR spectra
of organic free radicals, which often contain many lines resulting from the hyperfine
interaction with several H nuclei. Lines from interaction with C13 nuclei (I = 1/2,
natural abundance 1.11%) may also be observed; several examples are given by
Weil, Bolton, and Wertz [7]. However, whereas for free radicals and molecular
crystals the hyperfine lines in EPR spectra are usually resolved, this may not be
the case for paramagnetic centres in solids such as ionic crystals or semiconductors.
This occurs because in such materials there may be many more nuclei having a
SHF interaction with the paramagnetic centre and hence so many more lines that
they become superimposed resulting in the loss of resolution. It may, nevertheless,
be possible to measure the SHF interactions in such cases using the technique of
electron nuclear double resonance (ENDOR).
1312 D. Menard and R. Barklie

Ferromagnetic Resonance

FMR spectra typically look like a broad and high-intensity EPR single peak, with
comparatively strong anisotropy. In his initial experimental observation of FMR
absorption, Griffiths [5] reported resonance conditions significantly far from those
expected from the Larmor frequency. Kittel [15] is usually credited with the proper
explanation, highlighting the important role of the dipolar interaction leading to
strong shape anisotropy and demagnetizing effects. Kittel’s macroscopic approach
also includes effective fields associated with magnetocrystalline anisotropy.
As shown by Van Vleck, the microscopic quantum mechanical approach is,
within a reasonable approximation, equivalent to Kittel’s macroscopic phenomeno-
logical picture [16]. In his own words: “It is both an element of strength and
weakness in Kittel’s macroscopic method of handling the phenomenological energy
of anisotropy by means of an effective field that it frees one from the necessity of
specializing the model.” Also, the relaxation times involved in FMR are usually too
short to expect any significant Rabi oscillations of the longitudinal component of
the magnetization between the up and down states at a frequency of ω1 = |γ |B1 , as
discussed in Sect. “Microscopic Equation of Motion”. A small-response linearized
solution of (7) is usually fully justified.
It is common practice nowadays, in FMR studies, to use the Landau-Lifshitz-
Gilbert (LLG) equation, which is a combination of (9) and (7). Assuming that
the Gilbert parameter α is a constant (9) implies that the magnetization remains
constant in amplitude (but not in direction), so it describes the relaxation of a rigid
magnetization. This macrospin-like behavior is to be expected in ferro- and ferri-
magnetic materials due to strong exchange coupling, which tends to lock the relative
orientation of the moments together. It is also consistent with a viscous damping
proportional to the resonance frequency, another feature generally observed in
FMR experiments. Quenching of orbital angular momentum, as described in Sect.
“Anisotropic Zeeman Interaction”, is also quite effective in 3D ferromagnets,
leading to g-factor value of approximately 2 or equivalently to a gyromagnetic
ratio of approximately 28 GHz/T. Due to small spin-orbit corrections, the g-factor
value is in fact typically between 2.1 and 2.2 for most 3d-based materials, and it is
considered isotropic.

Effective Magnetic Fields

The macroscopic method is similar to the use of the spin Hamiltonian, previously
described for EPR. The macroscopic ground state is first calculated by minimizing
the magnetic part of the macroscopic free energy of the spin system, in the presence
of a static magnetic field, followed by calculating the linear response of (7) to a
microwave field excitation. The magnetic free energy, assuming, for the moment, a
single ferromagnetic domain with average uniform magnetization M, is

Utot = UZ + UMS + UA + UME . (39)


26 Electron Paramagnetic and Ferromagnetic Resonance 1313

In (39), the first term is the Zeeman effect,

UZ = −M · B 0 , (40)

analogous to (19), where B0 is the applied static field. The second term,

μ0
UMS = M · N · M, (41)
2

is the magnetostatic energy arising from the dipolar interaction between the spins, in
which N is the demagnetizing tensor. In contrast with EPR, the dipolar interaction
between the spins leads to a much stronger and sample-shape-dependent magnetic
anisotropy.
The other terms in (39), which are not associated with Maxwellian fields like
the first two, are due to other spin-spin or spin-orbit interactions and lead to
effective magnetic anisotropy fields, discussed further below. Here we separate the
magnetoelastic anisotropy term, UME , arising from external stress from the intrinsic
magnetocrystalline anisotropy term, UA .
Minimizing Utot (ϕ , θ ) with respect to the rigid magnetization orientation leads
to its equilibrium position, from which an effective field

∂U
B eff = − , (42)
∂M

can be calculated and used in (7), which is then solved assuming that the dynamic
parts of M and Beff are much smaller than their static parts (the linear approxima-
tion).
As a simplified example, consider an isotropic sample, magnetically saturated
along the z axis. Assuming B = B0 ẑ + μ0 he−iωt and M = M0 ẑ + me−iωt , where b
and m are small dynamic components of the fields perpendicular to z, (7) combined
with (9) leads to first order to a transverse susceptibility
    
mx χ −iκ hx
= , (43)
my iκ χ hy

where

ωM ω̃H ωM ω
χ= , κ= 2 , (44)
ω̃B − ω
2 2 ω̃B − ω2

and

ωM = |γ | μ0 M0 , (45)

ω̃B = ω0 + iαω, (46)


1314 D. Menard and R. Barklie

are in units of angular frequency. Equation (43) describes the response of the
dynamic magnetization to an internal dynamic field, exhibiting the characteristic
resonant response shown in Fig. 1, at the Larmor frequency ω0 . The main difference
with the isotropic EPR response of Sect. “Bloch Equations” is the use of a Gilbert
damping term rather than the Bloch term and the implicit condition mz = 0, arising
from the small signal approximation almost always justified in FMR. The use of
Gilbert damping term is formally equivalent to using a complex field as defined
by (46). Similarly, the use of the Bloch term would have resulted in a complex
frequency ω̃ = ω − i/T in (44).
The effect of magnetic anisotropy on the resonance condition, which is neglected
in (43–46), can be calculated using the Smith-Suhl formula [17]:

 
 2 2 1/2 
|γ | ∂ 2 Utot ∂ 2 Utot ∂ Utot 
ωres = −  , (47)
2 2 
M0 sin θ ∂θ ∂ϕ ∂θ ∂ϕ  θ = θ0
ϕ = ϕ0

in which the second derivatives are evaluated at the equilibrium position (θ 0 , ϕ0 ) of


the rigid magnetization, determined from

 
∂U  ∂U 
= = 0. (48)
∂θ θ=θ0 ∂ϕ ϕ=ϕ0

Self-consistent solutions of (47) and (48) generate a characteristic angle-


dependent FMR response relative to the rotation of the applied field around a
given crystallographic axis. Therefore, a combination of resonance field vs angle, at
a given frequency, around different rotation axes, enables one to precisely determine
the crystalline system and magnetic anisotropy constants. In situations where the
static equilibrium magnetization is oriented parallel to the applied field B0 , (48)
becomes trivial, but in general (47) and (48) must be solved iteratively to fit a set
of data. In what follows, we briefly highlight some contributions to the anisotropy
terms included in (39).

Shape Anisotropy

The dipolar interaction between spins is accounted for using (41), in which N is the
demagnetizing tensor with a unitary trace N xx + N yy + N zz = 1. It is a widespread
practice to consider a specimen of ellipsoidal shape, for which the demagnetizing
field, Bd = −μ0 N M, is uniform within the sample volume [18]. For fields applied
along one of the principal axes of the ellipsoid, N is diagonal. If the magnetically
saturating DC field is applied parallel to the z axis, the solution of (47) with energy
terms (40) and (41) yields the celebrated Kittel formula:
26 Electron Paramagnetic and Ferromagnetic Resonance 1315

  1/2
ωres = ω0 + Nyy − Nzz ωM [ω0 + (Nxx − Nzz ) ωM ] . (49)

Equation (49) shows that the resonance condition depends on the sample shape
and on the direction of the applied DC field relative to it. In ferromagnets, shape
anisotropy is quite significant, producing anisotropy fields on the order of μ0 M0 ,
which may be comparable to the applied fields. In paramagnets however, the shape
anisotropy fields are several orders of magnitude smaller because the magnetic
susceptibility is on the order of 10−5 .
Sample shapes, such as spheres, wires, and thin films, most usually encountered
in FMR experiments, can be considered as limiting cases of ellipsoids. Figure 5
illustrates the resonance condition as given by (49) for limited cases of ellipsoidal
samples, assuming their response is dominated by shape anisotropy. The values of
the demagnetizing factors for these cases are summarized in Table 1. Spherical sam-
ples exhibit no shape anisotropy, whereas thin films show large differences between
the responses to in-plane and out-of-plane applied fields. For non-ellipsoidal shapes,
the demagnetizing field is nonuniform, leading to line broadening and, sometimes,
to the excitation of nonuniform spin wave modes.

Fig. 5 Kittel’s resonance


condition for different (2)
samples dominated by shape
anisotropy. The arrows (3)
(1)
indicate the direction of the
saturating magnetic fields
(4)
Resonance frequency

(5)

Applied field

Table 1 Demagnetizing Configuration N xx N yy N zz


factors for sample shapes and
Cylinder parallel (1) 1/2 1/2 0
field directions shown in
Fig. 5 Thin film parallel (2) 0 1 0
Sphere (3) 1/3 1/3 1/3
Cylinder perpendicular (4) 1/2 0 1/2
Thin film perpendicular (5) 0 0 1
1316 D. Menard and R. Barklie

Magnetocrystalline Anisotropy

In crystalline samples, or in magnetostrictive amorphous samples with quenched-


in stress, a variety of spin-spin and spin-orbital interactions, including anisotropic
exchange, produce effective torques on the magnetization, leading to effective non-
Maxwellian fields in (7) or (47). These are accounted for using phenomenological
energy terms reflecting the symmetry of the interaction. For instance, cubic
anisotropy is described in (39) to second order with [12, 13]

UA = Kc1 αx2 αy2 + αy2 αz2 + αz2 αx2 + Kc2 αx2 αy2 αz2 , (50)

where Kc1 and Kc2 are first- and second-order anisotropy constants and α x , α y , α z
are the direction cosines of the magnetization with respect to the principal axes.
Similarly, uniaxial anisotropy leads to a term of the form [12, 13]

UA = Ku1 sin2 θ + Ku2 sin4 θ, (51)

where Ku1 and Ku2 are first- and second-order uniaxial anisotropy constants and θ
is the angle between the magnetization and the anisotropy axis. Formulae such as
(50) and (51) have been developed for all crystalline systems [19]. The anisotropy
field can also be expressed in terms of an effective demagnetizing tensor, analogous
to the demagnetizing tensor used in Sect. “Shape Anisotropy” [12, 13].
As the simplest example, let us consider a sphere with first-order uniaxial
anisotropy along the z axis, with a static field applied at an angle θ B from the z
axis, for which (47) predicts a resonance frequency [13]
  1/2
ωres = [ω0 cos (θ0 − θB ) + ωA cos 2θ0 ] ω0 cos (θ0 − θB ) + ωA cos2 θ0 ,
(52)

where ωA = |γ |2Ku1 /M0 is the effective uniaxial anisotropy field in units of angular
frequency. Further assuming that the applied field is sufficiently strong to align the
magnetization parallel to it, that is, θ 0 ∼
= θ B , which also implies ω0  ωA and an
equivalently high resonance frequency, (52) can be simplified into the following
expression for the angle-dependent resonance field:


∼ 1 1
B0 = ω + ωA 1 − 3cos θ .
2
(53)
|γ | 2

If the uniaxial anisotropy was strictly due to dipolar interaction in a wire


geometry, as treated in the previous section, then ωA = ωM /2, recovering cases
(1) and (4) of Fig. 5 when θ = 0 and π/2, respectively. Equation (53) is similar to
the fine structure EPR field (32), originating from dipolar interactions.
Magnetoelastic anisotropy may also be present in mechanically stressed materi-
als, such as thin films under epitaxial strain. A good example of magnetic anisotropy
dominated by the magnetoelastic term is rapidly solidified soft amorphous metal,
26 Electron Paramagnetic and Ferromagnetic Resonance 1317

such those as used in magnetoimpedance studies [20]. For instance, in soft


amorphous wires and ribbons, a first- order uniaxial term of the form

3
UME = λs σ sin2 θ (54)
2

is used to account for the extremely small anisotropy. In (54), σ is either a tensile
(σ > 0) or compressive (σ < 0) stress, defining the anisotropy axis, λs is the
magnetostriction constant (also positive or negative), and θ is the angle between the
magnetization and anisotropy axis. In these circumstances, (52) applies, in which
ωA = |γ |3λs σ /M0 , and FMR spectra enable the probing of mechanical stress or
magnetostriction if the other is known.
Finally, in FMR studies of ultrathin films and magnetic multilayers, a phe-
nomenological surface anisotropy is often included in the equations. This con-
tribution is usually highlighted in thin films of variable thicknesses, where the
surface anisotropy contribution appears inversely proportional to the thickness. A
comprehensive treatment of anisotropy in thin films is presented in Ref. [19].

Exchange Energy and Spin Wave Resonance

So far the exchange energy has been excluded from (47). Assuming isotropic
exchange, this term takes the form
 
μ0 nW 2 ∇M 2
Uex =− M +A , (55)
2 M0

where nW is the dimensionless Weiss molecular field constant and A is the exchange
stiffness (typically on the order of 10−11 J/m). Substituting the first term of
(55) into (42) leads to the Weiss effective field BW = μ0 nW M, which locks
the relative orientation of individual moments together into an average uniform
magnetization. While Bw can amount to several hundreds of Tesla, dominating
all other effective fields, it does not contribute to the torque in (7), as it is always
parallel to the magnetization. In addition, any expected anisotropy in the Weiss field
is effectively included in the magnetocrystalline anisotropy treated in the previous
section. The second term of (55), however, reflects the additional cost in energy due
to nonuniform magnetic excitations. Assuming a spin wave of the form m = m0 exp
(i k · r − iωt), (42) yields an effective field

bex = μ0 2ex k 2 m, (56)

where

2A
ex = (57)
μ0 M02
1318 D. Menard and R. Barklie

is the exchange length (typically a few nm). As a result, (46) generalizes to

ω̃B = ω0 + ωM 2ex k 2 + iαω, (58)

leading to the characteristic k2 dispersion associated with spin waves.


Assuming an unbounded isotropic ferromagnet, with the spin wave propagating
in a direction θ k relative to the equilibrium static magnetization, one can write the
dispersion relation as


1/2
ωres = ω0 + ωM 2ex k 2 ω0 + ωM 2ex k 2 + ωM sin2 θk . (59)

In this expression, the k2 terms are due to the short-range exchange interaction,
whereas the sin2 θ k arises from dynamic demagnetizing effects due to the
long-range dipolar interactions. Figure 6 illustrates this spin wave manifold for
unbounded isotropic materials. The shaded area between the two curves represents
the continuum of dipole-exchange spin waves that can be excited. For thin films or
slabs with finite dimension, the low k (long wavelength) spin waves are governed
by dipolar interactions and are influenced by the presence of surfaces. In these
circumstances, proper solutions of the electromagnetic problem, including boundary
conditions, yield a rich spectrum of magnetic excitations, including surface waves,
often called magnetostatic waves. Depending on the direction of the saturating field
relative to the surfaces, they are designated as backward volume waves, forward
volume waves, and Damon and Eshbach surface waves. A detailed treatment of the
subject can be found in Ref. [13].
Spin waves in magnetically ordered materials are usually excited by a nonuni-
form driving field, or they can arise due to pinned spin boundary conditions. They
occur naturally in bulk ferromagnetic metals, in which the skin effect results in a
driving field with gradually decreasing amplitude and phase shift as it penetrates the
material. They can be launched on purpose using localized microstrip antennae or
result accidentally from nonuniform demagnetizing fields or from the presence of

Fig. 6 Spin wave manifold w


in unbounded ferromagnets
angular frequency

wave vector k
26 Electron Paramagnetic and Ferromagnetic Resonance 1319

inhomogeneities in the material. A uniform driving field can also excite odd mode-
number standing spin waves in thin films with significant surface anisotropy. The
presence of spin waves in FMR spectra provides a mean to experimentally determine
the exchange length dex and exchange interaction A.

FMR in Metals

In principle, the behavior of ferromagnets submitted to a microwave field is the


solution of an electromagnetic boundary value problem, requiring the self-consistent
solution of Maxwell’s equations along with LLG equation. Maxwell’s equations,
with M as an explicit variable, B = μ0 (H + M), assuming Ohm’s law, J = σ E, a
scalar dielectric constant ε, and time harmonic fields ∼ exp (−iωt), combine into
the equation (after eliminating the E-field):

∇ 2 H − ∇ (∇ · M) + k02 (H + M) = 0, (60)

where k0 = (1 − i)/δ 0 , δ0 = 2/μ0 ωσ eff , and σ eff = σ + iωε is the an
effective conductivity including both ohmic and displacement currents. For good
conductors, k0 is complex and δ 0 is the nonmagnetic skin depth, whereas for

nonconductors, k0 = εμ0 ω is real. Assuming dynamic fields of the form m, h ∝
exp (i k · r − iωt), (60) can be written as
 
k02 I − kk
h= m, (61)
k 2 − k02
 
where k02 I − kk ij = k02 δij − ki kj is a second-order tensor. Equation (60) is the
dynamic part of the dipolar field. In addition, the linearized LLG equation leads to
a relation of the form h = X · m [13], which in the absence of anisotropy simplifies
to
 
1 ω0 + ωM 2ex k 2 + iαω iω
h= m. (62)
ωM − iω ω0 + ωM 2ex k 2 + iαω

The simultaneous solution (61) and (62) results in complex dispersion relations.
In the magnetostatic approximation, which is equivalent to setting k0 = 0 in (61),
one recovers the spin wave manifold (59) (provided damping is neglected setting
α = 0). Further setting k = 0 leads to the resonance conditions for uniform modes,
as treated above. However, these approximations are no longer justified in problems
involving conductive materials where the skin depth is smaller than their shorter
dimensions.
Admitting that k is perpendicular to the surface, due to skin effect in metals, the
local approximation, which consists of neglecting the k2 terms in (62), yields an
effective transverse permeability:
1320 D. Menard and R. Barklie

μeff ω2 − ω2 + 2iω ωar


= 2ar , (63)
μ0 ωres − ω2 + 2iω ωres

exhibiting ferromagnetic resonance at



ωres = ω0 (ω0 + ωM ), (64)

and ferromagnetic antiresonance at ωar = ω0 + ωM , with linewidths given,


respectively, by Δωres = α(2ω0 + ωM ) and Δωar = 2α(ω0 + ωM ). The inclusion
of magnetic anisotropy by adding appropriate terms in (62) poses no difficulties and
leads to a modification of ω0 in (64) and what follows.
Equation (64) shows that the FMR condition in bulk metals is identical to curve
(2) in Fig. 5, corresponding to thin films magnetized in-plane. This is due to the
dynamic demagnetizing effect, provided by the spin wave perpendicular to the
surface, arising from the skin effect. The antiresonance frequency corresponds to
a minimum of the magnetic permeability associated with a maximum penetration
depth (transparency) into the metal, quite similar to the plasma frequency associated
with dielectric permittivity.
In the strong skin effect regime, the electromagnetic waves in the metal are
characterized by an effective skin depth:

2
δeff = , (65)
μeff ωσ

in which the effective scalar permeability is given by (63). The FMR response is no
longer described in terms of complex magnetic susceptibility but rather in terms of
complex impedance [24]. Recognizing that the impedance in metals, in conditions
of strong skin effect, is inversely proportional to δ eff , (65) is indeed the source of
the giant magnetoimpedance effect (GMI) [20]. In the other limit of the weak skin
effect regime, for which δ eff is much larger than the thickness of the sample, one
may use the same formalism as for nonmetallic ferromagnets but with an additional
contribution to the damping due to eddy currents.
If we now relax the local approximation, putting the exchange term back into
(62), the dispersion relation becomes a fourth-order equation in k2 [21]. The four
pairs of roots correspond to two pairs of electromagnetic dispersion branches
and two pairs of spin waves, with strong hybridization around resonance. A
linear combination of the waves is necessary to satisfy the electromagnetic and
magnetization boundary conditions, such as free spin, pinned spin, or a mix of the
two [22]. The solution for the impedance finally appears as a 4 × 4 determinant
equation , which is relatively complicated, especially when anisotropy terms are
included in (62). Fortunately, approximations can be justified to simplify the result
considerably, with only small impact on the accuracy. A comprehensive analysis has
been presented [11, 23, 24].
26 Electron Paramagnetic and Ferromagnetic Resonance 1321

Experimental Observation of Electron Magnetic Resonance

Historically, EPR and FMR experiments have been performed with the same type
of spectrometer, using microwave cavities in which the frequency ω of a relatively
small B1 field is kept constant, while B0 is swept through the resonance at a rate slow
enough that relaxation maintains a constant magnetization in the rotating coordinate
system. In these conditions, the frequency ω1 is far less than the relaxation rate so
that the magnetization remains close to thermal equilibrium and exhibits a small
precession with approximately constant Mz around the average static effective
field. This classical setup has changed little in half a century, apart from some
improvement in the components for producing the microwave fields and digitally
acquiring the signals.
The basic instrumentation is presented in Sect. “Frequency Domain Cavity
EPR/FMR”. Alternative methods are then surveyed in Sect. “Survey of Other
Methods”. This second part is far from exhaustive as the wealth of EPR and FMR
methods would need another chapter, if not a book. The interested reader will find
an alternative complementary treatment in Ref. [25].

Frequency Domain Cavity EPR/FMR

Classical EPR/FMR spectrometers have been designed to operate over a range


of microwave frequencies from a few gigahertz up to hundreds of gigahertz, in
pulsed and continuous wave (CW) modes. Pulsed EPR, which will be discussed
in the next section, is less common than CW EPR. The use of a cavity mode
for CW operation implies that the frequency is fixed by the cavity and the field
is swept. The frequencies are associated with waveguide bands corresponding
to the frequency range for the single mode operation of the hollow metallic
waveguides used to carry the signals. Table 2 summarizes a few standard bands
frequently encountered in EPR/FMR measurements. Fields are typically provided
by a water-cooled electromagnet, except for the higher frequency bands that require
superconducting magnets. For reasons of suitability, convenience, and cost, the most
widely used EPR/FMR spectrometers operate at X-band (around 10 GHz).

Table 2 Standard frequency Band name Frequency range (GHz)


bands associated with
X 8.20–12.40
rectangular waveguides
Ku 12.40–18.00
K 18.00–26.50
Ka 26.50–40.00
Q 33.00–50.00
V 50.00–75.00
W 75.00–110.0
1322 D. Menard and R. Barklie

Fig. 7 Block diagram of a prototypical homodyne EPR spectrometer

Figure 7 illustrates the main components of the spectrometer. The microwave


source, traditionally a klystron or Gunn diode, has increasingly been replaced by
a broadband solid- state synthesizer. The signal propagates in rectangular hollow
metallic waveguides or coaxial cables, impinges on a resonator loaded with the
sample, is transmitted or reflected, depending on the type of resonator, and is
detected at the receiver. In the case of a reflection resonator, a circulator or a coupler
is used to redirect the signal to the receiver. Several secondary components, such as
isolators to protect the source from reflections and attenuators to adjust the power,
are not shown in the figure.
The microwave resonator is loaded with the sample. It amplifies the B1 driving
field and allows for efficient conversion of the sample response into a variation of
the reflected microwave signal. The Q factor is defined as Q = ωres /Δω, where ωres
is the angular resonance frequency of the resonator and Δω is its resonance width
at half power. High Q resonators traditionally consist of metallic cavities, coupled
to waveguides through iris holes or to waveguides with small antennae, with typical
values of Q between 500 and 20000 depending on the application. The use of low Q
coplanar waveguide or microstrip resonators is widespread in modern systems using
broadband techniques and will be discussed in the next section.
The frequency of the source is adjusted to excite a standing-wave mode in
the cavity. The sample is placed at a maximum of the microwave magnetic field.
Typically, the spectrometer is operated with the cavity critically coupled to the input
so that, in the absence of an EPR signal and for a non-lossy sample, all the incident
power is dissipated in the walls of the cavity. As these walls are usually highly
conductive so that the loss per cycle is low (and the cavity Q value high), the
26 Electron Paramagnetic and Ferromagnetic Resonance 1323

fields within the cavity build up to values greatly exceeding those of the incident
microwaves.
The coupling to the cavity can be adjusted with a metal-tipped screw just in
front of the cavity iris. When power is absorbed in the sample during resonance, the
cavity is no longer critically coupled, and some microwave power is reflected back
via the circulator, or transmitted, to a detector diode. In order to be most sensitive
to this reflected power, the diode must be appropriately biased. This is achieved by
passing some of the source output to the diode via the reference arm. Adjustment of
the phase allows either the absorption or dispersion mode of the EPR signal to be
detected.
As the external DC field is swept through resonance, the Q factor of the cavity
changes, and its resonance frequency is shifted, associated with the absorption and
the dispersion of the magnetic susceptibility, respectively. The frequency of the
source is usually locked onto that of the cavity so that only absorption is observed
and the reflected signal monitors changes in the Q factor of the cavity. In absorption
mode, Q exhibits a dip as a function of the sweeping field with the minimum at
the resonance field. The automatic frequency control (AFC) can be achieved using
frequency modulation of the source. This results in an amplitude modulation at the
diode, yielding the correction signal to adjust the source frequency to that of the
cavity.
It is also customary to superimpose a small field modulation on the static field.
This modulation is typically at 100 kHz, and its amplitude should be chosen to be
less than the signal linewidth. The diode output then contains an AC component at
100 kHz with an amplitude proportional to the slope of the bell-shaped resonance
curve. This AC component becomes the signal input to the phase-sensitive detector
(PSD) whose reference input comes from the oscillator driving the modulation coils.
The DC output of the PSD is a measure of the amplitude of its AC input so that the
EPR signal represents the derivative with respect to field of the absorption curve.
There are two advantages of this technique: at 100 kHz, the 1/f noise of the detector
is greatly reduced compared with its value at the sweep frequency, and the noise is
reduced because the PSD acts as a narrow band-pass filter centred on the modulation
frequency.
The signal amplitude S, which is inversely proportional to the minimum
detectable number Nmin of paramagnetic centres, depends on several parameters:

 1 B0
S∝ PMW × Bmod × Q × η × √ × , (66)
kB TD ν B1/2

where PMW is the microwave power incident on the cavity; Bmod is the amplitude of
the modulation field; Q has its value for the cavity loaded with the sample; η is the
filling factor, which is a measure of the fraction of microwave energy that interacts
with the sample; TD is the temperature of the detector; Δν is the bandwidth of the
detection system; ΔB1/2 is the half-width at half-height of the single absorption line;
and B0 is the resonance field.
1324 D. Menard and R. Barklie

The signal can be increased both by increasing the modulation amplitude and
microwave power, but if the former is too high, distortion will occur and for the
latter saturation can set in. Increasing the sample size will also boost the signal, but
there is little point in extending the sample beyond the region in the cavity where the
microwave field is concentrated. Working at higher frequencies, in smaller cavities,
helps to increase the sensitivity as it means a larger B0 and, if the sample volume
is constant, also a larger η. Finally, the noise can be reduced by signal averaging
over several sweeps or by taking longer measurements. It is frequently the value of
ΔB1/2 that limits the signal amplitude. For typical operating values, Nmin is in the
range 1011 –1012 .
There are several variants of the setup shown in Fig. 7, such as using microwave
amplitude modulation instead of field modulation, using a transmission cavity rather
than reflection, using mixers for up-conversion and down-conversion of the signals
instead of diodes, sampling the signal with a digitizer rather than conditioning it with
a lock-in detector, and so on. Classical setups are reviewed in detail by Poole [26].

Survey of Other Methods

Frequency Domain Techniques


Frequency domain techniques require low Q broadband resonators. They allow
for frequency swept FMR over a large frequency bandwidth and do not use any
automatic frequency control. Frequency sweep at fixed field is usually faster than
field sweep and often more convenient when high magnetic fields are generated by
a superconducting magnet. A fairly strong signal is also necessary, as the low Q
resonator has low sensitivity. Thus, these methods are mostly used for FMR rather
than EPR.
A straightforward approach to implement a broadband measure into a classical
EPR/FMR system is to replace the cavity with a shorted waveguide and disconnect
the AFC in the setup of Fig. 7. A modern version consists in using coaxial lines
instead of waveguides, along with striplines or coplanar waveguide resonators [27].
In contrast with limited bandwidth waveguides (Table 2), coaxial cables have no
lower cutoff frequency, and some of them can support frequencies from DC to
65 GHz. Over the last two decades, vector network analyzers (VNA) have become
the preferred instrument to implement these techniques. VNAs are versatile and
sensitive instruments, integrating both the excitation and detection stages of the
spectrometer.
Figure 8 illustrates such a broadband FMR setup, in which the VNA acts as
the source and detector, including most of the signal conditioning elements.The
resonator is a segment of a transmission line, here in a coplanar waveguide
(CPW) configuration, in which the magnetic sample is included as part of its
dielectric component. Stripline resonators are also exploited in these systems,
but CPW circuits have the additional benefit that they can be contacted using
microwave probes. Field modulation, sometimes used, is optional with a VNA.
Similar approaches were also developed in the late 1990s, for broadband GMI
26 Electron Paramagnetic and Ferromagnetic Resonance 1325

Fig. 8 Block diagram of a


prototypical VNA FMR
spectrometer. The sample is
placed on the CPW resonator,
modifying the dielectric
response of the transmission
line segment

Fig. 9 (a) Illustration (not to scale) of the CoFeB nanowire array built into a microwave
transmission line, sitting on a ground plane with a stripline on top, (b) measured hysteresis loop,
and (c) real and imaginary parts of the effective permeability at several applied fields corresponding
to different saturated or remanent states of the wire system. Details in [29] (Reprinted figure with
permission from [29]. Copyright (2009) by the American Physical Society)

studies, but in which the soft amorphous ferromagnetic metallic wire is part of the
conductive line the transmission line, which is very convenient to carry on FMR or
GMI from a few MHz up to several GHz [28].
Figure 9 shows an example of FMR data collected using a broadband VNA-based
method. The sample consisted in an array of ferromagnetic nanowires electroplated
in nanoporous alumina templates.The resonator was a stripline circuit fabricated
around the magnetic sample, the latter being part of the effective dielectric of
the device under test, in between the conductive lines. Both the dispersive and
absorptive parts of the complex susceptibility could be extracted as a function of
frequency, for different applied fields [29].
1326 D. Menard and R. Barklie

Time-Domain Techniques
Progress in high-speed electronics has also permitted time-domain pulsed EPR
experiments, such as free induction decay or spin echoes, gradually catching up
with extremely well-developed pulsed NMR techniques carried out at much lower
(radio) frequencies [10]. The method is to keep B0 fixed, while a sequence of short
high-power pulses of B1 field is applied, such that the Rabi frequency ω1 is higher
than the relaxation rate. The aim is to prepare the magnetization in a state out of
equilibrium and then measure its evolution by monitoring its recovery to the thermal
equilibrium state. Since T1 and T2 are nearly always much shorter for the electronic
than for the nuclear moments, the pulses used for EPR must be much shorter (ns)
and stronger (kW) than those used in NMR.
Compared to a CW spectrometer, a pulsed EPR spectrometer requires a pulse
shaping stage followed by high-power traveling-wave tube amplifier (TWTA) in
the excitation arm, providing the pulse sequence to a broadband low Q resonator.
A precise control over the timing and synchronization of the pulse shaping,
amplification, and data collection is critical. Figure 10 shows the main components
of a pulsed EPR spectrometer with a transmission resonator. Various components,
such as isolators, attenuators, and filters, are not shown for simplicity. The arbitrary
waveform generator (AWG) allows for pulse shaping, frequency up-conversion
(optional), and synchronization of the different components, including the PIN
diodes (mw switches) for protection of the receiver. Down-conversion of the signal
using a mixer enables the probing of the evolution of the magnetization in the
rotating frame so that the Fourier transform of the time response can display the
resonance peaks around a central frequency. Modern equipment permit fast digital
sampling or subsampling of the signal.

f
Attenuator Phase
Power
divider shifter

Mixer PIN PIN LNA Mixer


Diode TWTA Diode Signal conditionning
MW Source
and data acquisition

Resonator with
Pulse shaping magnetic sample
and frequency
upshift (AWG)
Electromagnets

Fig. 10 Simplified block diagram of a pulsed EPR spectrometer. The AWG control the timing
and synchronization of the components as well as the pulse shaping of the CW signal. Various
components, such as attenuators, isolators, and filters, are not shown for simplicity.
26 Electron Paramagnetic and Ferromagnetic Resonance 1327

Time-domain measurement is more challenging for FMR than it is for EPR,


due to the comparatively higher relaxation rate of magnetically ordered materi-
als. A noncoherent version of pulsed EPR, called pulsed inductive microwave
magnetometry (PIMM), has nevertheless been developed. It exploits a coplanar
waveguide resonator, such as shown in Fig. 8, to probe the free induction decay of a
ferromagnetic sample driven out of equilibrium by a short magnetic pulse [30]. The
Fourier transform of the time response is then used to extract the FMR spectrum at
each measurement field. Results are consistent with those obtained from broadband
FMR [27].

Electrically Detected Magnetic Resonance


By electrically detected magnetic resonance (EDMR) , we mean any variations of
the contactless EPR/FMR methods discussed previously but with the addition of
an electrical contact to probe a signal related to the resonance. EDMR traditionally
involved the rectification of a microwave signal in a magnetoresistive conductor
[31]. As the sample is submitted to a microwave field, an electrical current is
induced, which mixes with the electrical resistance, that is modulated at the driving
frequency by the precession of the magnetization. The mixing of the current and
resistance results in a dc voltage, which is proportional to the precession angle,
exhibiting the characteristic magnetic resonance response as the dc field is swept.
The method is limited to conductive samples and requires electrical contact to the
sample in order to extract the voltage.
A variation on EDMR exploits the spin Hall effect (SHE) in bilayers consisting of
ferromagnets coupled to a heavy transition metal, such as Pt or Pd. The experimental
setup is similar to that of the traditional EDMR, where the microwave signal
results in an electrical spin current diffusing in the heavy metal, which is then
transduced into a dc voltage thanks to the asymmetrical spin orbit diffusion of
spin up and spin down electrons [25, 32, 33]. As in the rectified EDMR, the
voltage exhibits the characteristic FMR response as the field is swept. While a
bilayer can simultaneously exhibit SHE and rectification, the distinct symmetry of
the two responses in angle-resolved measurements allows us to discriminate them
[32, 33]. There is an excellent review of the topic, discussing various experimental
configurations to implement EDMR [33].
More generally, the recognition of a spin torque effect in giant magnetoresistive
multilayered devices has opened up several new avenues for probing FMR. These
spintronic-based approaches exploit the conversion of a dc current into a microwave
response and vice versa. There are several variations of these methods, depend-
ing on the combination of dc and/or microwave excitation and detection in the
magnetoresistive device. Figure 11 shows one such example, where a multilayered
magnetoresistive device is submitted to a dc current and a microwave signal
generated from the sample is amplified by a broadband low- noise amplifier (LNA)
and displayed on a spectrum analyzer. The microwave signal may be due to spin
torque or simply a consequence of thermal noise. Indeed, thermal fluctuations
of the magnetization, which result in the modulation of the magnetoresistance,
submitted to a dc current, should produce a voltage spectrum reflecting the larger
1328 D. Menard and R. Barklie

Fig. 11 Block diagram of a Bias Tee


setup used to detect FMR Spectrum
generated by a dc current, analyzer
either through a spin torque LNA
effect or due to thermal
magnetic noise
DC Source

Sample interfaced with


broadband mw probe Electromagnets

susceptibility of the magnetization to fluctuate at natural resonances, in agreement


with the fluctuation dissipation theorem. This approach is sometimes referred to
thermal magnetic noise spectroscopy.

Optically Detected Magnetic Resonance


The use of IR or visible light to excite or detect microwave magnetic resonances
offers several other opportunities to extend the performance of classical EPR/FMR
setups. The physical mechanisms to bridge microwave and optical photons include
the magneto-optic response of the sample, the optical pumping of spin-polarized
carriers, and the laser-induced excitation of thermal magnons. In optically detected
magnetic resonance (ODMR), the resonances can be detected via a change in
intensity or polarization of an optical probe transmitted through or reflected from the
magnetic sample or from a change in luminescence intensity. In many cases, optical
detection leads to more sensitive or faster measurement than classical EPR/FMR
approaches.
For instance, time-resolved magneto-optic Kerr effect (TR-MOKE) is used to
probe the free induction decay in time-domain pulsed FMR experiments. In these
methods, the precession of the magnetization modulates the off-diagonal component
of the dielectric tensor, leading to a slight modulation of the light polarization inci-
dent on the sample. The reflected light is detected using stroboscopic subsampling
. This permits the probing of magnetization dynamics above 100 GHz, increasingly
challenging with microwave-only approaches as the frequency increases. An exam-
ple of such a setup is described in Ref. [25].
Another strategy lies into the combination of optical pumping and probing, in
which the intensity (photon counts) of the optical signal depends sensitively on
the spin population of optically pumped energy levels, which are controlled by
sweeping the magnetic field or the microwave frequency. This has been exploited in
systems with paramagnetic defects, such as nitrogen-vacancy centres in diamond,
for developing a new class of sensitive magnetometers. The same setup can be
used to detect ENDOR, in which NMR transitions between SHF split levels are
detected via a change in the optical signal (a change in luminescence intensity
or polarization), rather than by detecting the change in the EPR signal intensity,
26 Electron Paramagnetic and Ferromagnetic Resonance 1329

as in commercial ENDOR-capable EPR systems. An excellent discussion of this


technique together with that of optically detected EPR is given by Spaeth, Niklas,
and Bartram [9].

Conclusion

The observation of EPR or FMR in condensed matter systems provides information


on their magnetic properties, microscopic structure, and chemical composition. Both
methods are associated with electron spin resonance, sharing the same physical
principles, and therefore they are described by almost the same equation of motion.
What differentiates EPR from FMR is mostly the nature and magnitude of the
interactions experienced by their elementary magnetic moments, as phenomeno-
logically included in the effective fields in (7). This usually produces signals with
significantly different anisotropy and damping for the two techniques.
While the practitioners of the two methods historically adopted their own
formalisms and language to treat these interactions, EPR microscopically in terms
of effective spin Hamiltonian and FMR macroscopically in terms of free energy,
they are indeed very similar. Both approaches use phenomenological parameters
in a mathematical form consistent with the symmetry of the interactions. The
parameters, which can be calculated theoretically in principle, are often determined
experimentally by fitting the angle-dependent data with the calculated phenomeno-
logical response.
Most of the experimental methods capable of probing EPR and FMR require
the material under test to be submitted to external magnetic fields and microwave
radiation. The classical EPR/FMR spectrometer developed several decades ago is
still widely used today. However, new development in fast electronics, photonics,
and spintronics keeps extending the possibilities of these powerful characterization
techniques.

References
1. Purcell, E.M., Torrey, H.C., Pound, R.V.: Resonance absorption by nuclear magnetic moments
in a solid. Phys. Rev. 69(1–2), 37–38 (1946)
2. Bloch, F., Hansen, W., Packard, M.: Nuclear induction. Phys. Rev. 69(3–4), 127 (1946)
3. Zavoisky, E.K.: Spin magnetic resonance in the decimetre-wave region. J. Phys. USSR. 10,
197–198 (1946)
4. Cummerow, R.L., Halliday, D.: Paramagnetic losses in two Manganous salts. Phys. Rev.
70(5–6), 433 (1946)
5. Griffiths, J.H.E.: Anomalous high-frequency resistance of ferromagnetic metals. Nature.
158(670–671) (1946)
6. Abragam, A., Bleaney, B.: Electron Paramagnetic Resonance of Transition Ions. Clarendon,
Oxford (1970)
7. Weil, J.A., Bolton, J.R., Wertz, J.E.: Electron Paramagnetic Resonance: Elementary Theory
and Practical Applications. Wiley, New York (1994)
1330 D. Menard and R. Barklie

8. Pake, G.E., Estle, T.L.: The Physical Principles of Electron Paramagnetic Resonance. Ben-
jamin, New York (1973)
9. Spaeth, J.-M., Niklas, J.R., Bartram, R.H.: Structural Analysis of Point Defects in Solids:
An Introduction to Multiple Magnetic Resonance Spectroscopy. Springer-Verlag, Berlin,
Heidelberg (1992)
10. Slichter, C.P.: Principles of Magnetic Resonance. Springer-Verlag, Berlin Heidelberg (1990)
11. Vonsovskii, S.V.: Ferromagnetic Resonance: The Phenomenon of Resonant Absorption of a
High-Frequency Magnetic Field in Ferromagnetic Substances. Pergamon Press Ltd., Oxford
(1966)
12. Gurevich, A.G., Melkov, G.A.: Magnetization Oscillations and Waves. CRC Press (1996)
13. Stancil, D.D., Prabhakar, A.: Spin Waves: Theory and Applications. Springer (2009)
14. Bloch, F.: Nuclear induction. Phys. Rev. 70(7–8), 460–474 (1946)
15. Kittel, C.: On the theory of ferromagnetic resonance absorption. Phys. Rev. 73(2), 155–161
(1948)
16. Van Vleck, J.H.: Concerning the theory of ferromagnetic resonance absorption. Phys. Rev.
78(3), 266–274 (1950)
17. Smit, J., Belgers, H.G.: Ferromagnetic resonance absorption in BaFe12 O19 , a high anisotropy
crystal. Philips Res. Rep. 10(2), 113–130 (1955)
18. Osborn, J.A.: Demagnetizing factors of the general ellipsoid. Phys. Rev. 67(11–12), 351–357
(1945)
19. Farle, M.: Ferromagnetic resonance of ultrathin metallic layers. Rep. Prog. Phys. 61, 755–826
(1998)
20. Knobel, M., Vázquez, M., Kraus, L.: Giant magnetoimpedance. In: Buschow, K.H.J. (ed.)
Handbook of Magnetic Materials, vol. 15. Elsevier, London (2003)
21. Ament, W.S., Rado, G.T.: Electromagnetic effects of spin wave resonance in ferromagnetic
metals. Phys. Rev. 97(6), 1558–1566 (1955)
22. Rado, G.T., Weertman, J.R.: Spin-wave resonance in a ferromagnetic metal. J. Phys. Chem.
Solids. 11(3–4), 315–333 (1959)
23. Patton, C.E.: Classical theory of spin-wave dispersion for ferromagnetic metals. Czechoslov. J.
Phys.., Sect. B. 26(8), 925–935 (1976)
24. Frait, Z., Fraitova, D.: In: Romanov, A.S.B., Sinha, S.K. (eds.) Spin Waves and Magnetic
Excitations. Elsevier, Amsterdam (1988)
25. Farle, M., Silva, T., Woltersdorf, G.: Spin dynamics in the time and frequency domain. In:
Zabel, H., Farle, M. (eds.) Magnetic Nanostructures: Spin Dynamics and Spin Transport.
Springer (2013)
26. Poole, C.P.: Electron Spin Resonance: A Comprehensive Treatise on Experimental Techniques,
Second Edition. Dover (1983)
27. Kalarickal, S.S., Krivosik, P., Wu, M., Patton, C.E., Schneider, M.L., Kabos, P., Silva,
T.J., Nibarger, J.P.: Ferromagnetic resonance linewidth in metallic thin films: comparison of
measurement methods. J. Appl. Phys. 99(9), 093909 (2006)
28. Ménard, D., Britel, M., Ciureanu, P., Yelon, A.: High frequency impedance spectra of soft
amorphous fibers. J. Appl. Phys. 81(8), 4032–4034 (1997)
29. Boucher, V., Carignan, L.-P., Kodera, T., Caloz, C., Yelon, A., Ménard, D.: Effective perme-
ability tensor and double resonance of interacting bistable ferromagnetic nanowires. Phys. Rev.
B. 80(22), 224402 (2009)
30. Kos, A.B., Silva, T.J., Kabos, P.: Pulsed inductive microwave magnetometer. Rev. Sci. Instrum.
73(10), 3563–3569 (2002)
31. Egan, W.G., Juretschke, H.J.: DC detection of ferromagnetic resonance in thin nickel films. J.
Appl. Phys. 34(5), 1477–1484 (1963)
32. Azevedo, A., Vilela-Leão, L.H., Rodríguez-Suárez, R.L., Lacerda Santos, A.F., Rezende, S.M.:
Spin pumping and anisotropic magnetoresistance voltages in magnetic bilayers: theory and
experiment. Phys. Rev. B. 83(14), 144402 (2011)
33. Harder, M., Gui, W., Hu, C.-M.: Electrical detection of magnetization dynamics via spin
rectification effects. Phys. Rep. 661, 1–59 (2016)
26 Electron Paramagnetic and Ferromagnetic Resonance 1331

Bibliography
Abragam, A., Bleaney, B.: Electron Paramagnetic Resonance of Transition Ions. Clarendon,
Oxford (1970)
Weil, J.A., Bolton, J.R., Wertz, J.E.: Electron Paramagnetic Resonance: Elementary Theory and
Practical Applications. Wiley, New York (1994)
Pake, G.E., Estle, T.L.: The Physical Principles of Electron Paramagnetic Resonance. Benjamin,
New York (1973)
Spaeth, J.-M., Niklas, J.R., Bartram, R.H.: Structural Analysis of Point Defects in Solids: An
Introduction to Multiple Magnetic Resonance Spectroscopy. Springer-Verlag, Berlin, Heidelberg
(1992)
Slichter, C.P.: Principles of Magnetic Resonance. Springer-Verlag, Berlin Heidelberg (1990)
Vonsovskii, S.V.: Ferromagnetic Resonance: The Phenomenon of Resonant Absorption of a High-
Frequency Magnetic Field in Ferromagnetic Substances. Pergamon Press Ltd., Oxford (1966)
Gurevich, A.G., Melkov, G.A.: Magnetization Oscillations and Waves. CRC Press (1996)
Stancil, D.D., Prabhakar, A.: Spin Waves: Theory and Applications. Springer (2009)

David Menard received his Ph.D. from Polytechnique Montréal


in 1999. After a Postdoctoral Fellowship at Colorado State
University, he worked for Seagate Technology, USA, on giant
magnetoresistive heads. He joined Polytechnique Montreal, in
2003, where he established a research group in basic and applied
magnetics. His research activities focus on fundamental and prac-
tical questions of magnetic sensing and magnetization dynamics.

Robert Barklie obtained a PhD in Physics from Cambridge


University in 1969. He is now an Emeritus Fellow at Trinity
College, Dublin having retired in 2008 after 34 years lecturing
and researching in Physics. His main research interests are the
application of electron magnetic resonance techniques to the
study of paramagnetic centres in semiconductors and insulators.
Magnetization Dynamics
27
Andrew D. Kent , Hendrik Ohldag , Hermann A. Dürr,
and Jonathan Z. Sun

Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1334
Spin-Transfer Torque, Magnetic Switching, and Oscillations . . . . . . . . . . . . . . . . . . . . . . . . 1337
Spin-Transfer Torque in Phenomenological Form . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1337
Spin-Torque-Driven Anti-damping Magnetic Switching . . . . . . . . . . . . . . . . . . . . . . . . . . 1339
Orthogonal Spin-Torque-Driven Magnetic Switching . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1342
Spin-Torque Oscillators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1344
Spin-Transfer-Induced Excitation of Spin Waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1344
Spin-Transfer Vortex Oscillators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1346
Synchrotron and Femtosecond Laser-Based Time-Resolved Spin Dynamics . . . . . . . . . . . . 1347
Switching Schemes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1348
Microscopy Using Visible Light or X-Rays . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1349
Ultrafast Spin-Transfer Torques . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1351
Summary and Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1354
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1355

A. D. Kent
Center for Quantum Phenomena, Department of Physics, New York University, New York,
NY, USA
e-mail: andy.kent@nyu.edu
H. Ohldag
Advanced Light Source, Lawrence Berkeley National Laboratory, Berkeley, CA, USA
Department of Physics, University of California Santa Cruz, Santa Cruz, CA, USA
Department of Materials Science, Stanford University, Stanford, CA, USA
e-mail: hohldag@lbl.gov
H. A. Dürr
Department of Physics and Astronomy, Uppsala University, Uppsala, Sweden
e-mail: hermann.durr@physics.uu.se
J. Z. Sun ()
IBM T. J. Watson Research Center, Yorktown Heights, NY, USA
e-mail: jonsun@us.ibm.com

© Springer Nature Switzerland AG 2021 1333


J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_27
1334 A. D. Kent et al.

Abstract

Magnetism primarily describes the physics and materials science of systems


presenting a magnetization – a macroscopic order parameter characterizing
electron angular momentum. The order parameter is associated with the elec-
tronic exchange interactions, which are fundamentally quantum mechanical. Its
dynamic behavior bridges the macroscopic and the microscopic worlds. On
macroscopic length and time scales, it interacts with electromagnetic fields
dictated by Maxwell’s equations. On a microscopic scale, it involves the
quantum-mechanical electronic states both in spin space and momentum space,
thus giving rise to a wide range of behaviors that extend down to femtoseconds.
Thanks to the development of modern metrology, there have been many new
and noteworthy observations of magnetism-related phenomena across the entire
range – from spin-torque-induced anti-damping dynamics to ultrafast laser-
induced femtosecond electron dynamics that involve spin current and angular
momentum conservation. In this review, we introduce some observations on
magnetodynamics and the scientific subjects these new results give rise to.

Introduction

Magnetodynamics spans the entire electromagnetic spectrum. At low energies


and low frequency, the macroscopic dipolar interaction between a ferromagnetic
moment and the long-range electromagnetic force is the foundation for all modern
electromagnetic power conversion technology. These typically involve time scales
greater than a millisecond where electromagnetic induction and its interaction
with mechanical forces are the primary concern. At shorter time scales, between
a millisecond and a nanosecond, one enters the so-called micromagnetic regime,
which characterizes the collective motion of magnetic moments under an externally
applied magnetic field and their magnetic anisotropy field. These are still relatively
low-energy processes. Their characteristic time scale is of order τ0 ∼ 1/(γ μ0 Heff )
where Heff is the effective magnetic field, typically of the order of the saturation
magnetization of a ferromagnet M, corresponding to a few Bohr magnetons per
unit cell, and γ is the gyromagnetic ratio. μ0 Heff is typically several Teslas,
which gives a characteristic spin precession frequency of γ μ0 Heff of ∼28 GHz.
Long-range interactions between spins are usually dominated by magnetic dipole
interactions. On length scales below a few hundred nanometers and time scales
of a nanosecond or less, the exchange interaction between neighboring magnetic
moments becomes more significant. However, for magnetization dynamics on larger
length scales, moments still behave collectively and essentially in a classical manner.
In this regime, from long-range dipolar interactions to long-wavelength exchange
energy (“Long-wavelength exchange energy” in reference to low-energy spin-
wave modes (magnons) far from zone boundaries), the magnetodynamics can be
represented by the phenomenological Landau-Lifshitz-Gilbert, or LLG, equation
[1, 2]:
27 Magnetization Dynamics 1335

dM α dM
= −γ μ0 M × Heff + M× (1)
dt |M| dt

where γ = |g| μB /h̄ is the magnitude of the gyromagnetic ratio, g  −2, M is the
magnetization, and Heff is the effective magnetic field, including applied and dipolar
fields, and includes derivatives of M if dimensions approach that of the exchange
length, as discussed in Eq. 2. Here α is a dimensionless phenomenological quantity
describing damping – analogous to mechanical friction or the inverse of a Q-factor
used in describing a classical resonator.
√ In the limit of magnet size small compared to the exchange length [3] λex 
A/Ku where A is the exchange constant and Ku is the anisotropy energy density,
the system can be approximately viewed as having a single macrospin moment
described by two degrees of freedoms – i.e., its rotation angles.
Micromagnetics based on the LLG equation remains the most practical method-
ology for understanding dynamics in applications today – applications such as
for thin films in magnetic storage (hard disk and tapes) as well as in solid-state
integrated magnetic random access memories (MRAMs).
For time scales on the order of picoseconds or shorter, with energies above
tens of millivolts per unit cell, the atomic nature of the electronic spin becomes
a significant factor. This is the regime where the electronic band structure and
its spin-space degrees of freedom entangle. These interactions lie at the heart of
ferromagnetism. The two most important interactions are (1) Heisenberg exchange,
originating from the fermionic nature of the electrons that dictates the occupancy
of its atomic orbital states and related band structures, and (2) a usually smaller
spin-orbit interaction, which also plays a role, particularly in materials with heavy
elements. This is the regime where electronic transport processes gets coupled to
that of electron spin degrees of freedom. It gives rise to such properties as the
magnetoresistance and anomalous Hall effect in ferromagnetic conductors, where
the magnetic configuration affects the electrical transport.
Over the last two decades, we began to understand that there also exists an
inverse process – namely, that a transport electronic current carried by itinerant
electrons with spins can affect the magnetic state of the conductor as well. For
magnetodynamics, this interaction with transport current adds new terms to the
LLG equation which describe the so-called spin-torque effects. Such effects couple
magnetodynamics and electronic transport current beyond the dipolar-inductive
interaction. These understandings highlight the importance of angular momentum
conservation in the detailed balance of spin-polarized electronic transport.
The recent ability to control magnetization on ultrafast time scales (pico- or
femtoseconds) and short length scales (nanometers) have attracted interest as such
processes may play an important role for information storage and processing
technologies in the future. For these, classical approaches to describe such systems
are often insufficient, and it is necessary to describe and characterize the magnetic
properties of these physical systems on an atomic scale.
For time scales much shorter than a picosecond, most of the time, one would
need a quantum-mechanical picture of electronic excitation. The combined electron
1336 A. D. Kent et al.

Fig. 1 Energy reservoirs in a


ferromagnet. Exchange
between these is limited by
fundamental time constants.
Electron-phonon relaxation
times are around 1–10 ps,
spin-lattice relaxation times
are of the order of 100 ps, and
spin-orbit relaxation times are
of the order of 100 fs

spin system can often be modeled by an atomistic approach [4], which uses an
LLG equation as a semi-classical approximation on an atomic scale, with spins of
fixed length that are exchange coupled via a Heisenberg term. Such models can
numerically describe fast dynamics such as ultrafast laser-induced demagnetization.
An intuitive way to understand excitations involving magnetism is to view the
system as composed of a few “energy reservoirs” with characteristic coupling
strengths (i.e., time scales), as illustrated in Fig. 1. While this as often is an
oversimplification, it highlights the important factors at play and the approximate
time and energy for the various leading order factors. In principle, it is possible
to excite any of these reservoirs through pulsed or periodic optic, magnetic, or
electric fields. For example, the spin-orbit coupling energy is typically a few tens of
meV. If one applies a very strong and short field pulse that is able to overcome this
energy barrier on an atomic scale, it will only take about 100 fs for spin and orbital
momentum to realign, the spin-orbit relaxation time. Effect on these time scales
becomes relevant when the magnetization is excited using, for example, ultrafast
laser pulses that are able to perturb the electron system. However, in cases where
external magnetic fields or current pulses are applied, the typical electron-phonon
relaxation time of 1–10 ps and spin-lattice relaxation time of typically ≈100 ps play
a more important role since such excitations are limited to frequencies of a few tens
of GHz compared to optical laser excitations.
In what follows, we discuss several recent topics in magnetodynamics that
become important, thanks both to the advances of fundamental understanding and
to the significant development of modern metrology in X-ray and laser pump-
probe measurement technologies. We start with a discussion on the phenomena of
spin angular momentum transfer and the basic nonlinear dynamics involved, which
affects ferromagnetic systems in the form of a new term in the LLG equation.
The presence of spin-torque brings us a new, spin current-induced switching
mechanism for nanostructured magnets. It also brings a rich variety of nonlinear
27 Magnetization Dynamics 1337

magnetodynamics that have been most directly probed by the recently developed
X-ray spin-resolved microscopy. These new discoveries highlight the importance of
spin angular momentum current. We will also briefly review the dynamics occurring
on shorter time scales, often investigated by using femtosecond laser excitations. A
more extensive review on the femtosecond-level magnetic excitation can be found
in a separate chapter by Theo Rasing et al. in this volume.

Spin-Transfer Torque, Magnetic Switching, and Oscillations

Spin-transfer torque (STT) is a term that describes the dephasing effect of a spin-
polarized current upon entering a magnetically ordered material at a well-defined
interface. The dephasing of the spin component transverse to the magnetization
under the usually strong s-d exchange-like interaction results in a net torque being
exerted on the ferromagnetic moment. The effect was first proposed in spin-valve-
like structures [5, 6] as a manifestation for a type of energy-nonconserving torque
discussed in a ferromagnetic tunnel junction [7]. Such an energy-nonconserving
torque acts along the same vector axis as the damping torque. With the right
orientation of the spin polarization and the spin current direction, the STT can
overcome the damping torque in a nanomagnet and result in a magnetic reversal
if the nanomagnet is situated in an uniaxial anisotropy potential [5, 8, 9, 10]. The
same torque can also, under more general energy landscapes for the magnetic
moment, promote magnetic instability that results in persistent magnetic precession
or emission of spin waves [6, 11, 12].

Spin-Transfer Torque in Phenomenological Form

Spin current induced magnetodynamics can in most situations be treated approxi-


mately as a torque exerted on the magnetic moment. This is due to a separation of
time scales – the time scale involved in conduction electron spin’s dephasing is of
the order of s-d exchange energy (of the order 0.1 – 1 eV, ∼1 fs) and is generally
much shorter (of higher energy scale by several orders of magnitudes) than the time
scale involved in typical ferromagnetic material’s magnetodynamics in a few Tesla
(of the order of 0.1 ns) of combined anisotropy, dipolar, and external magnetic fields.
Phenomenologically and in macrospin limit, the STT can be described as an
additional term to the Landau-Lifshitz-Gilbert equation, in the form of
     
1 dnm α dnm Is
= μ0 H eff × nm + nm × − nm × (nm × ns ) , (2)
γ dt γ dt γm

with nm = m/m as the time-dependent magnetic moment’s unit vector direction,


H eff the total effective magnetic field on the moment (including anisotropy, dipolar
demagnetization, and applied field), α the LLG damping coefficient, and Is the spin
current, here written with a practical unit of magnetic moment/second (Am2 /s), and
1338 A. D. Kent et al.

H
damping
I < Ic
I > Ic
q spin-torque I >> Ic
(a) M (b)

Fig. 2 An illustration of the dynamics involved in spin-torque-induced magnetic switching. (a)


The vector torque relationship. The dephasing-induced spin-torque has the same axial direction as
the damping torque, in the direction of nm × (nm × ns ) as described in Eq. 2. It can be parallel or
antiparallel to the damping torque depending on the sign of the spin current. (b) The trajectory of
a macrospin on its unit sphere under the combined influence of damping and spin-torque. When
the spin current is less than the threshold Ic necessary for net zero damping, the moment damps
toward its north-pole energy minimum. When spin current exceeds threshold Ic , an anti-damping
precession ensues in the northern hemisphere, with cone angle θ increasing over time. The rate of
cone angle opening is faster for larger spin current

ns the spin polarization unit vector direction (Fig. 2). γ = |g| μB /h̄ is the magnitude
of the gyromagnetic ratio constant. μB here is the Bohr magneton, and g  −2 for
simple ferromagnetic metals is the Landé g-factor. For a charge current Icg with spin
polarization factor of 0 < η < 1, for example, Is = (μB /e) Icg η.
When the magnetic body in volume Ω is larger than the relevant magnetic length
and the system is not in its macrospin limit, the quantities in Eq. 2 take on a position
dependence in addition to the time dependence. Then Eq. 2 becomes a constitutive
equation that, with appropriate boundary conditions, determines the micromagnetic
dynamics of the system, with the usual non-local dipolar field interactions described
by Hd r, r  . The location dependence of the magnetic moment also introduces an
exchange energy-related term into Eq. 2 [13]. These result in a modification of the
field term as
  
  A 2
H eff → H eff (r) + Hd r, r  d 3 r  + nm (r) , (3)
Ω μ0 Ms

which describes the long-wavelength effect of ferromagnetic exchange, with A


being the magnetic exchange energy, and a replacement in Eq. 2 of a macrospin
direction vector with a unit vector for the direction of the local saturation magneti-
zation, that is, m → Ms = Ms nm (r).
The source of interface spin current Is may include spin filtering of a spin-
polarized current in metals, such as the Valet-Fert type of current-perpendicular-
27 Magnetization Dynamics 1339

to-plane (CPP) spin-valve structures, either diffusive or ballistic [14, 15]. It may
originate from spin filtering through a barrier in magnetic tunnel junctions (MTJs)
[16, 17]. The spin current could also be generated by transport processes involving
a strong spin-orbit interaction [18, 19, 20, 21, 22, 23]. Large spin-torques have also
been reported at interfaces between topological insulators and ferromagnets; see
Refs. [24,25,26], for example, for a review on the subject. The spin current can also
be induced by heat flow involving a magnon thermal gradient [27, 28, 29]. For these
non-spin-filtering mechanisms, the electron to spin current conversion is in principle
not limited to one h̄/2 per electron charge current, whereas for the spin-filtering
processes, one electron can only transfer one h̄/2 of spin angular momentum, that
is, one Bohr magneton μB of magnetic moment, which sets a limit to the charge-spin
conversion efficiency.
Two factors make STT dynamics rich and often complex. They are (1) the
effect of finite temperature, which makes thermal fluctuations and thus related
stochastic distribution issues important for STT-driven dynamics, and (2), in most
realistic samples, various internal degrees of magnetic freedom that participate in
the STT dynamics, making the whole process more complex than a macrospin.
Both are significant scientific challenges to one’s ability to quantitatively understand
and control the STT-related behavior. Both are critically important for technology
development.
The primary effect of spin-torque on a magnetic nanostructure is to change its
effective magnetic damping. When a sufficient amount of spin-torque is applied in
the right direction, the net damping of the nanomagnet can become negative for a
certain region on the macrospin coordinate unit sphere, which can amplify magnetic
precession (see Fig. 2). For a nanomagnet with uniaxial magnetic anisotropy, this
can result in its moment reversing from the easy-axis direction anti-aligned to the
spin polarization to the direction aligned with it [5, 8, 10]. In other geometries, it
is possible to send the moment into persistent oscillation mode, which becomes an
STT-driven oscillator [30].

Spin-Torque-Driven Anti-damping Magnetic Switching

Besides its novel anti-damping magnetodynamics, STT-driven magnetic switching


is technologically interesting primarily because it can effectively manipulate and
switch a nanomagnet situated in an anisotropy potential. A nanomagnet in a uniaxial
anisotropy potential with at least two stable moment direction states is the building
block of a magnetic memory bit. The STT mechanism provides a local current-
based addressing mechanism, circumventing the long-range nature of magnetic
field-based write mechanisms such as a magnetic write head, thus enabling high-
density, all-solid-state magnetic memory (STT-MRAM).
For a macrospin, from Eq. 2, the instability threshold for a collinearly aligned
anisotropy axis, applied field direction, and spin current polarization direction, the
threshold switching spin current is simply
1340 A. D. Kent et al.

  
2e α
Ic0 = μ0 mHeff , (4)
h̄ η

where η is the effective spin polarization of the charge current, whose exact
functional form depends on the structure of the nanomagnet’s spin current transport
environment [5, 31, 17, 32, 33]. Heff = Hk + Ha in the case of a simple uniaxial
anisotropy with its anisotropy field as Hk and the applied field Ha in the same
direction.
In the low-temperature T → 0 K limit, a current larger than Ic0 is required to
initiate the anti-damping switching. The time it takes to switch a nanomagnet in the
macrospin limit is inversely proportional to I − Ic0 in the form of
 
τ0 π
τ≈ ln , (5)
I /Ic0 − 1 2θ0

with θ0 being the initial angle the moment makes with its easy axis upon the
m/μB 1
switching on of the spin-torque and τ0 = = [10, 34, 33].
ηIc0 /e αγ μ0 (Hk + Ha )
At finite temperature, there is a thermal distribution of initial angles θ0 about
the easy axis. The precessional reversal process would also be thermally scattered.
Of these two processes, in the short-time, high STT drive limit (strictly speaking
in the asymptotic limit of I /Ic0  10 [34]) where τ 1/(αγ μ0 Hk ), the initial
condition’s thermal distribution usually dominates the consideration [34,35]. Taking
this distribution into account, one arrives at the probability for switching at time
t for I  Ic0 as P (t), which has the high-barrier Eb  kB T and long-time
t  τ0 / (I /Ic0 − 1) asymptote [36, 37, 34, 38]
    
π 2 Eb − τ2t (I /Ic0 −1) π 2 Eb
P (t) ≈ exp − e 0 + O exp − , (6)
4kB T 4kB T

describing the P (t) → 1 limit.


In the sub-threshold I < Ic0 region, while no switching would be expected
at zero temperature, with finite temperature, thermally assisted reversal has a
finite probability. This switching probability, however small it might be, will also
be magnified by the application of spin-torque, and the resulting sub-threshold
switching probability in the macrospin limit is given by
     

Eb H ν1 I ν2
P (t) ≈ 1 − exp −f0 t exp − 1− 1− , (7)
kB T Hk Ic0

for 0  P (t) 1 with I Ic0 and H Hk , with f0 being the attempt frequency
for escape from the metastable magnetic state. ν1  2 if H is nearly collinear
with the anisotropy axis; otherwise ν1 < 2 [39], and 1 ≤ ν2 ≤ 2 depending on
the details of the anisotropy potential shape [40, 41], with ν2 → 2 for a perfectly
collinear, uniaxial anisotropy-only configuration. The exact values of the exponents
27 Magnetization Dynamics 1341

ν1,2 in an experimental nanomagnet structure (such as a magnetic tunnel junction)


depend further on details of the system’s micromagnetic behavior, as the dynamics
is usually not macrospin [42, 43, 44, 45, 46]. The exponents in such experimental
systems are generally in the range described above with some variations from device
to device, materials combinations, as well as different measurement time scales and
drive amplitudes.
Equations 6–7 form a pair of asymptotic descriptions of the probabilistic
switching behavior of a macrospin under collinear spin-torque. Note that neither
covers accurately the “most-likely switching” region of I ∼ Ic0 or when P ∼ 1/2.
For this crossover region’s accurate mathematical macrospin-solution, a full Fokker-
Planck equation-based numerical treatment is often necessary [47, 37, 38].
Experimentally, the STT-induced magnetic switch was observed earlier in highly
spin-polarized manganite tunnel junctions [8], in nanostructured spin valves [9],
and in magnetic tunnel junctions (MTJs) [16]. The advent of MTJs, especially
later the MgO barrier-based MTJs [48, 49, 50, 51], with perpendicular magnetic
anisotropy [52, 53, 54] significantly accelerated technological development in
memory applications. This is due to their impedance match into existing CMOS
circuits and their large magnetoresistance (MR) for readout. Importantly, spin-
torque switching of PMA-based magnetic storage elements can be efficient, i.e.,
they have a high ratio of the thermal stability (Eb ) of the element to the switching
current (Ic0 ), much higher than in-plane magnetized elements (where the threshold
current goes as the easy-plane anisotropy, which can be large  M, but the
thermal stability is set by the easy-axis anisotropy, which is typically M).
The perpendicular MTJ element diameter can also be scaled to very small sizes
(with 11 nm diameter demonstrated on CMOS integrated wafers [55] and 8 nm on
individual junctions [56] at the time of this writing).
Early experimental observations of STT-driven switches were made over a
time scale typically longer than a μs. These at finite temperature only probe
the sub-threshold, thermal activation-mediated switching process. The drive speed
dependence of such switching, either with STT or with field drive (easy-axis
hysteresis measuring Hc ’s rate dependence), can be used to estimate the thermal
activation barrier height Eb [57, 58, 59, 44].
Nanosecond-level super-threshold switching was first observed in metallic spin
valves, demonstrating the characteristic switching speed and drive amplitude trade-
off similar to Eq. 5 [60, 61, 62, 63, 64]. A similar fast-switching characteristic
was also observed in MTJ-based switching events on back-end CMOS integrated
devices [59]. Nanosecond speed STT-induced switching using non-local spin
current was also seen in spin-valve-like filtered structures [65] and in spin Hall-
induced spin currents [66, 67].
Fast, reliable switching is essential for memory applications of STT-based
devices. To this end, a significant amount of effort is directed to experimentally
assessing the switching statistics as it depends on the drive voltage’s amplitude and
duration in a PMA MTJ. For macrospins in a uniaxial anisotropy potential with
barrier height Eb , a relatively simple relationship is obtained between the driving
spin current amplitude, the drive duration, and the amount of switching error ( r in
1342 A. D. Kent et al.

the parameter Q0 , defined below), in the asymptotic form of

Eb Q0
Iw ≈ + , (8)
κ τw

for a write pulse width in the range of τw ∼ τ0 or smaller. Here, Iw is the write
current amplitude. κ = Eb /Ic0 ≈ (h̄/4e) (η/α) is the so-called STT “switching
efficiency” here expressed in the macrospin limit [44, 45]. From Eq. 6 and Ref. [44]
    
e m π 2 Eb
Q0 ≈ ln , (9)
2η μB 4kB T r

with r = 1 − P (τw ) representing the write error probability. Q0 /e is the number


of electrons Q0 /e = (Iw − Ic0 )τw needed to switch the magnetization; for a
current spin polarization η ∼ 1, Q0 /e is the order of the number of elemental spins
(Bohr magnetons) in the free layer. Thus Eq. 8 expresses the conservation of spin
angular momentum in spin-transfer switching. While the derivation of Eq. 8 relies
on a macrospin model, all parameters involved (Eb , κ, Q0 , τw , and Ic0 ) can be
determined in experiment. The resulting Iw is to the leading order a viable estimate
even beyond the macrospin limit. The departure from macrospin is mainly captured
by sub-volume nucleation of thermally activated reversal states [44, 68, 69], which
makes the value of κ size dependent and below its macrospin value, and by Q0
being drive amplitude dependent and smaller than macrospin value [70].
STT-switched PMA MTJs were demonstrated to switch at 10 ns pulse width
reliably with a switching error below 10−11 per switching operation on single device
level [71]. The most recently demonstrated 10 ns switch with deep statistics was on
an 11 -nm-diameter PMA MTJ at an error rate of 10−9 [55]. Even at 11 nm diameter,
the device’s behavior is still not completely in the macrospin limit but shows steeper
decrease of the switching error rate upon the increase of pulse height, indicative of
fractional volume initiation of the switching process [55, 44].
In this aspect of the STT device operation, one challenge for a successful
technological deployment of STT-based memory is to obtain high reliability STT
switching (with r 10−9 at the very least) with as small a value of Iw as possible
while retaining write speed and a sufficient thermal activation barrier height Eb
to ensure a 10-year nonvolatility of stored data bit. A review of the development
status of STT-MRAM and a comparison of STT-MRAM to other solid-state memory
technologies can be found in in Refs. [72, 70].

Orthogonal Spin-Torque-Driven Magnetic Switching

In addition to the anti-damping type of magnetic switching, when sufficiently


strong, a spin-torque can result in precessional switching. This type of switching
requires a spin-torque approximately 1/α stronger than the anti-damping type
of switching, since now the torque needs to drive dynamics associated with the
27 Magnetization Dynamics 1343

Fig. 3 Precessional orbits of


a biaxial macrospin. The red
lines show in-plane orbits
about the easy axis (x), while
the blue lines illustrate
out-of-plane magnetic orbits
about the hard axis (z). The
black line is the separatrix,
marking the boundary
between in-plane and
out-of-plane orbits. (Figure
adapted from Ref. [81])

magnetic anisotropy. This mechanism was originally proposed as a high-speed


precessional switching method [73]. It has also been recently demonstrated with
a slight modification [74, 75] in spin Hall effect-based orthogonal spin-torque
three-terminal type of devices [76, 74, 77, 75, 21, 78, 79, 80], where the spin Hall
effect-induced spin current was proved to be sufficiently strong, possibly with some
help from a “magnetic field-like” component as well due to symmetry breaking
spin-orbit interactions at the interfaces.
For an in-plane magnetized free layer, this requires spins with a polarization
component perpendicular to the plane. The free layer then has biaxial magnetic
anisotropy with an easy axis in the film plane (typically set by an asymmetric shape
of the element, e.g., an ellipse) and a hard axis perpendicular to the plane. There are
then two types of precessional orbits, as illustrated in Fig. 3 [82, 81]. One type is
around the easy axis and is similar to the orbits in a uniaxial nanomagnet discussed
above. The second type is orbits around the hard axis, orbits of the magnetization out
of the film plane. For a spin polarization collinear with the easy axis, the switching
is of the anti-damping type described above, and a spin-torque of the appropriate
sign (i.e., spin current polarity) results in a direct switching of the magnetization.
However, when there is a sufficient component of spin polarization out of the film
plane, out-of-plane orbits can be excited, and the switching can be precessional, with
a speed set by the demagnetization field (τ ∼ 1/(γ Ms )). This switching can be very
fast (<50 ps), but the switching probability will typically be an oscillatory function
of the pulse amplitude and duration, meaning that the pulse timing may be critical to
obtaining a low write error rate. Switching with current pulses as short as 500 ps has
been demonstrated in MTJ devices that incorporate a perpendicularly magnetized
spin-polarizing layer, known as an orthogonal spin-transfer (OST) device [83, 84],
and switching with even shorter current pulses has been reported in all metallic
1344 A. D. Kent et al.

spin-valve devices (a GMR-based device, in contrast to a MTJ) with a perpendicular


spin-polarizing layer [85,86,87,88]. In similar GMR structures, in low-temperature
experiments, direct evidence of coherent precessional magnetization switching with
clear oscillations in the switching probability were recently reported [89].
An exciting modeling result demonstrated that an adiabatically decaying current
pulse can lead to highly reliable precessional switching [90]. The essential reason
is that a spin-torque bias can be present, by a slight tilt of the spin polarization
direction away from the exact orthogonal direction of the anisotropy’s easy axis, as
the magnetization relaxes toward one of its easy directions, from the out-of-plane
orbits (blue curves in Fig. 3) to the in-plane orbits about the easy axis (red curves in
Fig. 3). As of this writing, this model remains to be tested in experiment.

Spin-Torque Oscillators

Spin-Transfer-Induced Excitation of Spin Waves

Spin-torque can cause persistent magnetic oscillations under many situations and
excite spin waves. This can either be of the form of coherent precession of a near
macrospin [10,30,91] or spatially non-uniform magnetization patterns [6,11,92,93,
94, 95, 96] of various wavelengths and amplitudes. In both cases, the magnetization
dynamics can be highly nonlinear.
A point contact to an otherwise extended ferromagnetic thin film stack is a
prototypical configuration that can be used to excite spin waves. The STT is
concentrated in the contact region and thus will generate a non-uniform magnetic
excitation (i.e., because regions outside the contact do not experience a significant
STT). A typical structure consists of a thin film stack in the form FM | NM | FM
(free) surface , and the point contact can be either via a metallic tip [92] or by
lithographic patterning [93] to create a “nanocontact,” an aperture with a diameter
200 nm in a dielectric that is filled with a non-magnetic metal (Fig. 4).
The magnetization excitations can be localized or extended. In the former case,
spin excitations are near the contact and decay rapidly outside the nanocontact
region, while in the latter case, propagating spin-wave modes can be excited. The
nature of the excited modes (extended or localized) depends on many characteristics
of the contacts, ferromagnetic layers, and applied current and field. The basic issue
is whether spins excited by STT in the contact region can couple to the propagating
spin waves in the free layer.
In a linear response region (limit of low currents, currents just above the
threshold for current-induced excitations), it was shown theoretically that the
wavelength of the spin waves would be approximately the nanocontact diameter
and propagating spin waves would be excited, with spin-wave amplitudes that decay
algebraically with distance from the nanocontact [97]. However, this situation is not
generic. At higher current amplitudes, spin waves can be localized in the contact
region [98, 99, 100], and the nature of the excitations depends on the applied
field (i.e., there can be transitions between localized and extended spin waves as
27 Magnetization Dynamics 1345

Fig. 4 Schematic of a lithographically defined nanocontact to a magnetic bilayer. Current flow in


the contact produces a STT on the moments in the contact region that excites spin waves. The spin
waves can propagate away from the contact under certain conditions

a function of the applied field [101]). Both extended and localized excitations have
been observed experimentally using optical and X-ray imaging methods, such as
Brillouin light scattering [102, 103], Kerr effect [104], and scanning transmission
X-ray microscopy [105, 106].
In the case in which the free layer has perpendicular magnetic anisotropy and
the field is perpendicular to the layer plane, the excitations are localized, and
spin waves can “condense” in the contact region forming a magnetic droplet
soliton [107], a state that was predicted theoretically in the 1970s for a ferromagnetic
layer with axial anisotropy and no dissipation or damping [108, 109]. This has
recently been observed experimentally [110, 111, 112], including directly using
scanning transmission X-ray microscopy (STXM) [105]. In the latter experiment,
the profile of the excitation in the contact region was directly measured, and the
spin-wave amplitude was seen to decay rapidly outside the contact, i.e., the spin
excitation was shown to be localized in the contact region. Modeling has shown
that STT in nanocontacts to thin films with perpendicular magnetic anisotropy
can also stabilize topological magnetic structures, including dynamical magnetic
skyrmions [113, 114]. Solitons were also directly observed in in-plane magnetized
free layers in which the Oersted field from the current leads to a more complex
potential for spin waves [106], for example, the potential favors propagation of spin
waves in particular directions away from the contact. In this case, STXM was used
to create movies of the magnetic excitations, using a heterodyne method of locking
the frequency of the spin excitations to an external microwave source. Further,
this directional propagation of spin waves has been used to phase-lock multiple
nanocontact oscillators [115].
1346 A. D. Kent et al.

The discovery of very large spin-orbit torques in heavy non-magnetic transition


metals (e.g., Pt, Ta, and W) [76, 21] has led to new types of spin-torque oscillators.
The geometry consists of a heavy metal film with an interface to a magnetic layer;
a prototype thin film structure is substrate | non-magnetic TM | FM (free) . In
these oscillators, the charge current flows in the plane of the non-magnetic transition
metal layer, but spin current flows perpendicular to the plane creating spin-wave
excitations in the free layer (or magnetic switching of the free layer as discussed
in section “Spin-Transfer Torque in Phenomenological Form”). Oscillators based
on few micron-diameter permalloy ferromagnetic disks [116, 117] and sub-micron
in-plane magnetized CoFeB elements [118] have been realized. In addition, various
contact and ferromagnet layer geometries have been demonstrated [119, 120, 121].
In addition, the phase locking of multiple spin Hall nano-oscillators was recently
demonstrated [122].
The fact that charge current only flows in the plane of the non-magnetic
transition metal and need not flow perpendicular to the plane (as in nanocontact- and
nanopillar-based spin-transfer devices) enables exciting magnetization dynamics
in magnetic insulators using spin-orbit torques. This is interesting for a number
of reasons, one of which is that spin waves can in principle travel over much
larger distances because magnetic insulators can have orders of magnitude lower
damping than conducting magnets. Another is that spin-orbit torques enable control
and modification of the spin-wave spectrum of magnetic insulators. A variety of
experimental studies of yttrium iron garnet (Y3 Fe5 O12 , YIG) films, a ferrimagnetic
insulator with very low damping (α < 10−4 ), have been reported recently showing
spin-orbit torque-induced persistent magnetization oscillations and the excitation
of propagating spin waves [123, 124, 125]. Spin excitations in magnetic insulators
using spin-transfer torques are new area of research that is developing rapidly at the
time of this writing.

Spin-Transfer Vortex Oscillators

Spin transfer can also be used to nucleate, excite, and study magnetic vortices and
can be applied to the study of other types of topological magnetic structures (e.g.,
skyrmions). Vortices can form in thin ferromagnetic layers and disks composed
of soft magnetic materials, such as permalloy. In a nanodisk, they minimize
magnetostatic energy at the cost of exchange and anisotropy energy, forming a
circular spin structure shown in Fig. 5. They can thus be favored by the geometry
of the magnetic layer; typically thicker disks will have a tendency to have vortex
ground-state magnetic configurations (see, for example, [126]). In a thin extended
layer, such as in a spin-transfer nanocontact (Fig. 5), vortices can be nucleated by
applying a current pulse, as the Oersted field from the pulse favors a circular spin
configuration [127]. The center of the vortex has a singular region, a core, in which
the moments point out of the film plane [128]. There are thus two physical quantities
that characterize a magnetic vortex, the sense of spin circulation in the plane (a
winding number) and the direction of core magnetization (a polarity). Both these
quantities play an important role in the dynamics of vortices – as described by
27 Magnetization Dynamics 1347

West East

SiO2
Free layer
Metal
Fixed layer
Current Bottom Current
electrode

Fig. 5 Schematic of a nanocontact with the free layer in a magnetic vortex configuration. A pulse
current can nucleate a vortex, the planar spin configuration shown above the contact, and currents
used to excite gyrotropic vortex motion. This motion can be studied by measuring the contact
resistance as a function of time or spectral measurements. (Figure adapted from Ref. [127])

the Thiele equation, an equation derived from the LLG equation that describes a
vortex’s “center of mass” motion as well as that of other types of magnetic objects
(e.g., domain walls) [129].
Part of the interest in vortex oscillators is that they can have high-quality
factors [130]. Their oscillation frequency is typically 0.1 to a few GHz, about
one order of magnitude less than spin-wave frequencies (i.e., the FMR frequency).
However, the oscillator’s output power can be large, particularly when the magnetic
layer in which the vortex forms is one of the electrodes of a magnetic tunnel junction
nanopillar [131]. Spin-torques also enabling exciting nonlinear dynamics and vortex
oscillators are a physics playground for exploring such dynamics, including chaotic
magnetization dynamics [132]. A variety of experimental studies have examined the
interactions between magnetic vortices, including in nanopillars with two magnetic
layers that each contains a vortex [133] and in proximal nanopillars [134].
Further, as with other types of spin-torque oscillators [135], injection locking of
vortex oscillators has been demonstrated [136]. Time-resolved imaging studies of
magnetic vortex dynamics have been conducted using MOKE [137] as well as X-ray
microscopy [138].

Synchrotron and Femtosecond Laser-Based Time-Resolved Spin


Dynamics

The previous sections focused on dynamics on the time scale of 10 to 100 s of


picoseconds. As described in the introduction, electrical transport measurements
are the workhorse for experimental characterization in this area. They make use
1348 A. D. Kent et al.

of magnetoresistance effects, and suitable electronics operating at tens of GHz can


provide information about the macroscopic dynamic magnetic behavior. However,
the exact microscopic behavior caused by the interplay of spin-torque with the
effective field and the exact shape of the element can only be indirectly deduced
if the transport results are compared to micromagnetic simulations. For direct
observation of many such dynamics involved in spin-torque-driven oscillations
in nanostructures, the recently developed spin-resolved X-ray microscopy proves
to be a useful tool. After describing the different switching schemes that are
used in X-ray-based and optical experiments, we will give a brief review on this
type of measurements. A more detailed review on this subject can be found in
Refs. [139, 140].

Switching Schemes

In the following, we are concerned with the manipulation and ultimately the
reversal of magnetization on fast time scales. Naively one would expect that
the magnetization of a sample reverses upon application of a magnetic field of
the opposite direction. However, this is not the case, since the origin of the
magnetization is the spin angular momentum of the atoms, and to change angular
momentum, one has to supply angular momentum to the sample. The most common
way to do this is it to apply a torque with an external field. That means that the
external field has to be applied at an angle to the magnetization. This is in general
described in the Landau-Lifshitz-Gilbert equation shown in Eq. 1, which includes
the gyromagnetic ratio γ , the effective field (sum of external field, demagnetization
field, and anisotropy fields), as well as the damping constant α, which is necessary
to include dissipative processes in a realistic medium.
Figure 6 shows three different possibilities to affect the magnetization by
supplying angular momentum or by applying a torque to the current magnetization
in an experiment. The panel on the left hand side shows the simple case of applying

Fig. 6 Illustration of different switching schemes, via application of a local magnetic field (left),
optical excitation (center), or spin-polarized current (right)
27 Magnetization Dynamics 1349

a magnetic field. While this approach is rather straightforward to realize and can be
described using Eq. 1, it has practical limitation given by the maximum field that
one can achieve and the fact that it is non-trivial to localize magnetic fields to the
area of interest. The fact that field-induced dynamics on larger length scales are
rather complex limits the speed with which the magnetization can be reversed using
external fields. Interestingly, one finds that even when using the extremely strong
fields of several Teslas generated by a relativistic electron bunch for a picosecond
or less, the magnetization far away from the excitation is still precessing for about
100 picosecond [141], essentially limiting the speed with which information stored
in such a system can be processed by only using conventional magnetic fields.
To speed up the reversal and to control the magnetization in a more confined
manner, other excitation mechanisms are considered. For example, the surprising
possibility to manipulate magnetization using short and powerful laser pulses has
evolved over the past decade [142, 143, 144, 145, 146]. Here a magnetic sample
is irradiated with an ultrashort laser pulse without the application of an external
field or current. The exact nature of the reversal mechanism and the role of the
angular momentum supplied by the photons are still debated. We will give a
short overview over this field in section “Ultrafast Spin-Transfer Torques”. A more
thorough overview will be given by T. Rasing et al. in another chapter.
One of the most efficient ways to detect magnetization dynamics on these
time scales is via electrical transport. This approach uses standard radiofrequency
analysis and lock-in equipment that is commercially available, which is why it is
widely used. For this detection method, one makes use of magnetoresistance effects
like giant magnetoresistance (GMR) or anisotropic magnetoresistance (AMR). Any
time one investigates a layered magnetic system, it is usually possible to use the
GMR effect, which manifests itself as a change in magnetization when a current
flows between two magnetic layers separated by a non-magnetic layer. This is
in particular useful for studies of spin-torque oscillators where the current flows
through a small nanocontact from a fixed polarizing layer to a free oscillating
layer. In this case, one can make use of the already existing electrical contacts and
directly determine the magnetization dynamics using a spectrum analyzer and lock-
in detection. If there is only one magnetic layer or the devices are simply planar,
one can use the AMR effect, which manifests itself in a change in resistance if the
magnetization is parallel or perpendicular to the current direction. This effect is at
least an order of magnitude smaller than the GMR, but it is compatible with most
experimental geometries. However, none of these approaches provide microscopic
or spatially resolved information. Typically, one can obtain indirect information
about the microscopic behavior of the sample by comparing the macroscopic
transport results to model calculations.

Microscopy Using Visible Light or X-Rays

In order to obtain microscopic information about magnetization dynamics, it is


possible to either use visible light microscopy or X-ray microscopy. Microscopy
1350 A. D. Kent et al.

Fig. 7 Two different approaches to image spin-wave dynamics excited by a nanocontact. The
image on the left is acquired using Brillouin light scattering from [102], while the image on the
right shows a localized spin wave acquired using scanning transmission X-ray microscopy [105]

using visible light based on the magneto-optical Kerr effect has the advantage that
the setup can be fairly simple and requires few components. Also using standard
pump-probe approaches, one can achieve very good time resolution of the order of
100 fs. However, the spatial resolution is limited, and the fact that the Kerr rotation is
generally small means that it is not generally possible to use time-resolved MOKE
microscopy. Nevertheless, using sophisticated lock-in approaches, one can today
use TR-MOKE to image different spin-wave modes on materials like YIG [147]
or permalloy [148]. Improved sensitivity to spin-wave dynamics in particular using
optical microscopy has been demonstrated more recently by using Brillouin light
scattering (BLS). In BLS, one images not the directly reflected light from the surface
of the magnetic sample but light scattered at a certain angle. The angle is determined
by the interaction of the incoming photon with magnons in the sample related to
the existence of magnetic excitations, like, e.g., spin waves. This approach provides
excellent sensitivity to magnetization dynamics, since there is very little background
at the observation angle, and in theory, all scattered light is related to the presence
of magnons. The left panel in Fig. 7 shows an image of the envelope of a spin wave
propagating from a nanocontact located underneath the triangular-shaped electrical
contact. Due to the limited penetration of the light, no information can be obtained
underneath the electrical contact.
One way to improve the spatial resolution and to gain the ability to study
buried magnetic layers is X-ray microscopy [149, 150], either in scanning or
full-field manner. These studies are usually conducted using a synchrotron X-ray
source, which provides short tunable and polarized X-ray pulses between 50 and
100 ps duration. The absorption of circular polarized X-rays at the L-absorption
resonances of the 3d transition metals like Fe, Co, and Ni between 500 and 1000 eV
depends strongly on the relative alignment of X-ray propagation and magneti-
zation (X-ray magnetic circular dichroism or XMCD), providing a suitable and
27 Magnetization Dynamics 1351

element-specific contrast mechanism for magnetic imaging. Due to the short wave-
length of such X-rays (1–10 nm), one is also able to obtain magnetic information on
the nanoscale [139]. If one combines X-ray microscopy with pump-probe schemes
and dedicated detection electronics [151, 152, 153], it is then possible to follow
magnetization dynamics with 10 nm spatial and 10 ps temporal resolution. One
example is shown on the right hand side of Fig. 7, where the presence of a magnetic
soliton generated by a spin-torque nano-oscillator is shown [105].
We note that another way to image magnetization dynamics is to use scan-
ning electron microscopy with polarization analysis (SEMPA); see, for example,
[154,155]. In SEMPA, a standard SEM instrument is combined with a Mott detector
to identify the spin of the emitted electrons. By using an orthogonal set of Mott
analyzers, it is then even possible to obtain a full three-dimensional map of the
magnetic behavior after excitation. However, due to the rather low efficiency of
the Mott detector, the very high surface sensitivity of the detector, and the vacuum
requirements, there are practical limitations on the types of samples that can be
studied.

Ultrafast Spin-Transfer Torques

Light is an important tool to manipulate and control magnetic properties. Advances


in the development of fs optical lasers enabled radically novel ways for probing
and controlling magnetism. At such sub-ps time scales, the ensuing non-adiabatic
dynamics is different from the LLG-based description in the rest of this chapter. It
will be reviewed in detail in chapter (T. Rasing) of this book. Briefly, the pioneering
observation of sub-picosecond demagnetization in ferromagnetic nickel following
fs optical laser excitation [156] was followed by the discovery of a wide range
of laser-induced phenomena in other metallic systems, ranging from the excitation
of precessional spin dynamics [157, 158, 159] to laser-induced magnetic phase
transitions [160, 161]. More recently deterministic so-called all-optical switching
by single femtosecond pulses of circularly polarized light in an increasing variety
of materials and heterostructures [162, 163, 164] has continued to intrigue and
stimulate researchers. These effects are based on manipulating magnetic interactions
such as exchange via the oscillating electric field of the exciting laser pulses. The
latter mainly acts on the electron charge, effectively heating the material’s electrons
far above the temperature of the lattice. This electron-light interaction, however,
conserves electron spin, so it is a central question how such charge excitations can
lead to a sub-ps collapse of magnetic order.
Here we describe recent developments utilizing the spin-conserving nature of
fs laser excitation of magnetic materials, i.e., laser-induced ultrafast spin currents,
to probe and understand their properties and attempt to utilize them for ultrafast
magnetic switching at a distance. Figure 8 displays the schematic excitation process
of electrons from occupied electronic states below the fermi level into unoccupied
states above. Theoretical modeling shows a superdiffusive spin transport of mainly
majority spins away from the excitation region [165]. Spin-conserving fs laser
1352 A. D. Kent et al.

Fig. 8 Schematic illustration


of the fs laser excitation of
superdiffusive spin currents in
a Ni film. The initially
ballistic hot electron motion
becomes superdiffusive via
scattering events altering
energy and momentum of the
excited electrons. The
different amount of scattering
for hot electrons with up and
down spins leads to an
effective spin current that can
escape into an adjacent layer
such as Al. (After Battiato
et al. [165])

excitation leads to an excited population of electrons with both up and down


spins shown in Fig. 8. Spin-dependent electronic scattering processes lead to longer
lifetimes for the majority spin (up) component. These spins can effectively travel a
longer distance and preferably leave the ferromagnetic layer generating a transient
spin polarization in an adjacent non-magnetic layer as shown in the figure.
Superdiffusive spin currents traversing a non-magnetic Au layer have been
detected via nonlinear second harmonic generation [166]. Ballistic Fe spins injected
into a Au layer traveling close to the Au Fermi velocity arrive at the Au back
interface within hundreds of femtoseconds, while a diffusive component was
detected at times up to 1 ps, in qualitative agreement with calculations [167]. Fe
layer thickness variations show that the active injection zone is a 1–2 -nm-thick
Fe layer at the Fe/Au interface [166, 168, 169]. This implies that superdiffusive
transport is of limited importance for ultrafast demagnetization of significantly
thicker ferromagnetic films, an observation supported by recent demagnetization
experiments in Ni films where no difference between front-side and back-side
pumping was observed [170]. References [171, 172] reported the detection of
superdiffusive spin currents in non-magnetic layers via the inverse spin Hall effect.
The use of materials displaying a large spin Hall effect allowed the detection of
superdiffusive spin currents via the characteristic emitted terahertz electromagnetic
pulse with a polarization determined by the transverse charge current in the spin
Hall layer [173, 174].
In layered structures with two ferromagnetic layers with collinear magnetic
moments separated by non-magnetic spacers, superdiffusive spin currents prefer-
entially excited in top magnetic layer have been found to affect the demagnetization
behavior in the bottom magnetic layer [175, 176, 177] (left panel of Fig. 9).
Replacing the spacer layer with an insulator was found to stop the superdiffusive
currents [175]. The role of transient spin accumulation at interfaces is currently
not well established. The longer ballistic mean free paths for majority spin (up)
electrons should lead to minority spin (down) accumulation at a ferromagnetic
27 Magnetization Dynamics 1353

Fig. 9 Schematic illustration of how superdiffusive spin currents generated by fs laser excitation
of a ferromagnetic layer (red bottom layer) can be utilized to manipulate spins in a (blue)
ferromagnetic layer separated by a non-magnetic spacer (yellow). (left) and (right) panels indicate
how superdiffusive spin currents can affect the magnetic moments of collinear spins and can lead
to a precessional spin motion, respectively. (From Hellman et al. [188])

interface layer upon injection of an unpolarized current from an adjacent non-


magnetic metal layer. However, reports for Au/Ni layers that this could even lead to
ultrafast demagnetization of 15 -nm-thick Ni films [178] remain controversial [179,
180]. However, a majority spin transmission through Pt(30 nm)/[Co/Pt] (6.5 nm)/Cu
(100–200 nm) heterostructures was observed via transient spin accumulation in
the Cu layer following fs laser heating of the Pt layer [181]. We finally point
out that the insertion of a tunneling barrier between two ferromagnetic layers
enables a new control mechanism via magnetic tunnel junctions. He et al. [182]
observed that spin tunneling through MgO spacers could influence the ultrafast
demagnetization of adjacent CoFeB layers. A first attempt at controlling the
femtosecond demagnetization in CoFeB-based magnetic tunnel junctions via tuning
the voltage applied to the junction was reported in [183].
One of the most typical spintronic devices is the spin-transfer torque magnetic
random access memory, where a spin current is used to exert a torque on a
magnetic bit ultimately switching its direction. Under typical operation conditions,
near-equilibrium, large-enough spin currents can only be generated in nanopillar
devices. The use of strong, ultrashort non-equilibrium spin currents could open
up new ways for spin-transfer torque switching. Schellekens et al. [184] and
Choi et al. [181] simultaneously demonstrated spin-torque-induced precession
dynamics driven by superdiffusive spin currents. These experiments utilize two
ferromagnetic layers with in-plane and out-of-plane magnetization separated by Cu
and Pt [184] and Cu [181] spacer layers of varying thickness. Although the induced
1354 A. D. Kent et al.

precession angles are still small due to the limited amount of observed spin angular
momentum transfer of several percent between the orthogonally magnetized layers,
such experiments represent a unique tool to actually quantify and consequently
optimize angular momentum transfer through interfaces.
A key ingredient in all-optical switching presented in more detail in the chapter
by Rasing et al. in this book is the dramatic reduction of the sample magne-
tization via ultrafast demagnetization. This allows angular momentum exchange
between magnetic subsystems to take over as clearly demonstrated by ultrafast
element-specific X-ray magnetometry in GdFeCo films [185]. The emergence of
a ferromagnetic transient state out of the antiferromagnetically aligned transition
metal and rare-earth magnetic subsystems bears this fingerprint of non-local angular
momentum exchange [185]. The question arises whether demagnetization accom-
panied by superdiffusive spin currents can also enable magnetic switching. Graves
et al. [186] found such a mechanism for GdFeCo using ultrafast X-ray scattering at
novel X-ray free-electron laser facilities. The basis of this study was the chemical
segregation observed in GdFeCo films into Gd-rich and Fe-rich nanoregions on a
10 nm length scale. Following ultrafast demagnetization of both the Gd 4f and Fe
3d magnetic sublattice, a reversal of the Gd 4f moments in Gd-rich nanoregions was
observed. This led to a local ferromagnetic state with parallel alignment of Fe 3d and
Gd 4f moments. The authors also measured the net spin angular momentum transfer
across ∼10 nm distances and showed that it is compatible with spins arising from
Fe 3d states. They must become delocalized from Fe-rich regions being transported
laterally into Gd-rich areas to accumulate there until the originally antiparallel Gd
4d magnetization is reversed [186]. These measurements lead the way to study the
effects of ultrafast spin transport over nanoscale dimensions. Although currently
only the action on local magnetic moments can be probed [186], future advances in
X-ray nano-spectroscopy offer the unique opportunity to also determine the effects
of non-local transport currents on local valence-level populations [187].

Summary and Outlook

Modern magnetism and magnetodynamics have seen significant progress in the last
two decades. The theoretical discovery and experimental observation of spin-torque
opened up new frontiers for fully localized control of nanomagnetic objects. This
provided a useful means of writing a magnetic bit in high-density arrangement such
as silicon-integrated STT-MRAM. Spin-torque also brought about a new type of
dynamics involving negative damping, opening up the world of spin current-driven
nonlinear magnetic oscillators, new ways of propagating and manipulating spin
waves, and new nonlinear magnetic condensates. Much of the nonlinear properties
of such magnetodynamics are still being explored at the time of this writing. For
such exploration, the new metrologies developed during the last two decades have
become very important; those of XMCD-based spin-resolved and time-resolved
microscopy with sub-20 nm resolution have enabled direct observation of many
such dynamic states, accelerating in many cases our understanding of such complex
27 Magnetization Dynamics 1355

systems. Beyond transport current-induced spin dynamics, light interaction with


magnetic materials has also been shown to induce significant non-equilibrium spin
angular momentum change of the electronic systems in such materials. These have
been demonstrated to result in ultrafast magnetic switching, with the electronic
process completing as fast as a few hundred femtoseconds. A deeper understanding
of the materials and electronic physics is well under way and will hopefully be
addressing future technology needs in the ultrafast realm. Like many other branches
of scientific inquiry, the field of magnetodynamics has been rejuvenated by an influx
of new ideas and new scientific methods as well as the development of ever-so-
more-advanced measurement capability. Whereas in the recent past such research
and development effort was often centered around magnetic storage and recording,
the frontier is now broader. Magnetism being fundamentally nonlinear and rich in
its dynamic behaviors continues to fascinate us all, both in revealing new scientific
insights and in providing new technologies for the future.

Acknowledgments A.D.K. acknowledges partial support by the National Science Foundation


under Award DMR-1610416. Work at NYU was also supported in part by the Quantum-Materials
for Energy Efficient Neuromorphic-Computing, an Energy Frontier Research Center funded by
DOE, Office of Science, BES, under Award DE-SC0019273.

References
1. Lifshitz, E.M., Pitaevskii, L.P.: Magnetism, Chapter 7. In: Statistical Physics, Part 2, A.
Wheaton & Co. Ltd., Exter II, 285 (1981). http://www.fulviofrisone.com/attachments/article/
208/Landau%20L.D.%20&%20Lifschitz%20E.M.-%20Vol.%209%20-%20Statistical
%20Physics%20part%202.pdf
2. Kittel, C.: Ferromagnetism and antiferromagnetism, Chapter 15. In: Introduction to Solid
State Physics, 6th edn., p 434. Wiley (1986)
3. Dennis, C.L., Borges, R.P., Buda, L.D., Ebels, U., Gregg, J.F., Hehn, M., Jouguelet, E.,
Ounadjela, K., Petej, I., Prejbeanu, I.L., Thornton, M.J.: The defining length scales of
mesomagnetism: a review. J. Phys. Condens. Matter 14, R1175 (2002)
4. Ellis, M.O.A., Evans, R.F.L., Ostler, T.A., Barker, J., Atxitia, U., Chubykalo-Fesenko, O.,
Chantrell, R.W.: The Landau-Lifshitz equation in atomistic models. Low Temp. Phys. 41,
705 (2015)
5. Slonczewski, J.C.: Current-driven excitation of magnetic multilayers. J. Magn. Magn. Mat.
159, L1 (1996)
6. Berger, L.: Emission of spin waves by a magnetic multilayer traversed by a current. Phys.
Rev. B 54, 9353 (1996)
7. Slonczewski, J.C.: Conductance and exchange coupling of two ferromagnets separated by a
tunneling barrier. Phys. Rev. B 39, 6995 (1989)
8. Sun, J.Z.: Current-driven magnetic switching in manganite trilayer junctions. J. Magn. Magn.
Mater. 202, 157 (1999)
9. Katine, J.A., Albert, F.J., Buhrman, R.A., Myers, E.B., Ralph, D.C.: Current-driven
magnetization reversal and spin-wave excitation in Co/Cu/Co pillars. Phys. Rev. Lett. 84,
3149 (2000)
10. Sun, J.Z.: Spin-current interaction with a monodomain magnetic body, a model study. Phys.
Rev. B 62, 570 (2000)
11. Slonczewski, J.C.: Excitation of spin waves by an electric current. J. Magn. Magn. Mater.
195, L261 (1999)
1356 A. D. Kent et al.

12. Bazaliy, Y.B., Jones, B.A., Zhang, S.-C.: Modification of the Landau-Lifshitz equation in the
presence of a spin-polarized current in colossal- and giant-magnetoresistive materials. Phys.
Rev. B 57, R3213 (1998)
13. Herring, C., Kittel, C.: On the theory of spin waves in ferromagnetic media. Phys. Rev. 81,
869 (1951)
14. Valet, T., Fert, A.: Theory of the perpendicular magnetoresistance in magnetic multilayers.
Phys. Rev. B 48, 7099 (1993)
15. Brataas, A., Nazarov, Y.V., Bauer, G.E.W.: Finite-element theory of transport in ferromagnet-
normal metal systems. Phys. Rev. Lett. 84, 2481 (2000)
16. Huai, Y., Albert, F., Nguyen, P., Pakala, M., Valet, T.: Observation of spin-transfer switching
in deep submicron-sized and low-resistance magnetic tunnel junctions. Appl. Phys. Lett. 84,
3118 (2004)
17. Slonczewski, J.C.: Currents, torques and polarization factors in magnetic tunnel junctions.
Phys. Rev. B 71, 024411 (2005)
18. Hirsch, J.E.: Spin Hall effect. Phys. Rev. Lett. 83, 1834 (1999)
19. Zhang, S.: Spin Hall effect in the presence of spin diffusion. Phys. Rev. Lett. 85, 393 (2000)
20. Kimura, T., Otani, Y., Sato, T., Takahashi, S., Maekawa, S.: Room-temperature reversible
spin Hall effect. Phys. Rev. Lett. 98, 156601 (2007)
21. Liu, L., Pai, C.-F., Li, Y., Tseng, H.W., Ralph, D.C., Buhrman, R.A.: Spin torque switching
with the giant spin hall effect of tantalum. Science 336, 555 (2012)
22. Kato, Y.K., Myers, R.C., Gossard, A.C., Awschalom, D.D.: Observation of the spin Hall
effect in semiconductors. Science 306, 1910 (2004)
23. Kajiwara, Y., Harii, K., Takahashi, S., Ohe, J., Uchida, K., Uchida, M., Mizuguchi, M.,
Umezawa, H., Kawai, H., Ando, K., Takanashi, K., Maekawa, S., Saitoh, E.: Transmission of
electrical signals by spin-wave interconversion in a magnetic insulator. Nature 464, 262–266
(2010)
24. Hasan, M.Z., Kane, C.L.: Topological insulators. Rev. Mod. Phys. 82, 3045 (2010)
25. Qi, X.-L., Zhang, S.-C.: Topological insulators and superconductors. Rev. Mod. Phys. 83,
1057 (2011)
26. Yan, B., Zhang, S.-C.: Topological materials. Rep. Prog. Phys. 75, 096501 (2012)
27. Slonczewski, J.C.: Initiation of spin-transfer torque by thermal transport from magnons. Phys.
Rev. B 82, 054403 (2010)
28. Nakata, K., Simon, P., Loss, D.: Wiedemann-Franz law for magnon transport. Phys. Rev. B
92, 134425 (2015)
29. Bauer, G.E.W., Saitoh, E., van Wees, B.J.: Spin caloritronics. Nat. Mater. 11, 391 (2012)
30. Kiselev, S.I., Sankey, J.C., Krivorotov, L.N., Emley, N.C., Schoelkopf, R.J., Buhrman, R.A.,
Ralph, D.C.: Microwave oscillations of a nanomagnet driven by a spin-polarized current.
Nature 425, 380 (2003)
31. Slonczewski, J.C.: The physics of spin-transfer in layered metallic magnets. EPFL Latsis
Symposium 02 extended abstract (2002)
32. Slonczewski, J.C.: Theory of spin-polarized current and spin-transfer torque in magnetic
multilayers. In: Kronmúller, H., Parkin, S. (eds.) Handbook of Magnetism and Advanced
Magnetic Materials. Spintronics and Magnetoelectronics, vol. 5. Wiley (2007). https://www.
wiley.com/en-us/Handbook+of+Magnetism+and+Advanced+Magnetic+Materials%2C+5+
Volume+Set-p-9780470022177
33. Sun, J.Z.: Physical principles of spin torque. In: Handbook of Spintronics. Springer Scientific
and Business Media, Dordrecht (2014). https://doi.org/10.1007/978-94-0077604-3-47-1
34. Liu, H., Bedau, D., Sun, J.Z., Mangin, S., Fullerton, E.E., Katine, J.A., Kent, A.D.: Dynamics
of spin torque switching in all-perpendicular spin-valve nanopillars. J. Magn. Magn. Mater.
358–359, 233 (2014)
35. Liu, H., Bedau, D., Sun, J.Z., Mangin, S., Fullerton, E.E., Katine, J.A., Kent, A.D.: Time-
resolved magnetic relaxation of a nanomagnet on subnanosecond time scales. Phys. Rev. B
85, 220405(R) (2012)
27 Magnetization Dynamics 1357

36. Sun, J.Z.: Write-efficiency and read-disturbance in a spin-torque switched nanomagnet. IBM
Internal Memo (2006)
37. He, J., Sun, J.Z., Zhang, S.: Switching speed distribution of spin-torque-induced magnetic
reversal. J. Appl. Phys. 101, 09A501 (2007)
38. Butler, W.H., Mewes, T., Mewes, C.K.A., Visscher, P.B., Rippard, W.H., Russek, S.E., Heindl,
R.: Switching distributions for perpendicular spin-torque devices within the macrospin
approximation. IEEE Trans. Magn. 48, 4684 (2012)
39. Coffey, W.T., Crothers, D.S.F., Dormann, J.L., Geoghegan, L.J., Kalmykov, Y.P., Waldron,
J.T., Wickstead, A.W.: Effect of an oblique magnetic field on the superparamagnetic
relaxation time. Phys. Rev. B 52, 15951 (1995)
40. Pinna, D., Kent, A.D., Stein, D.L.: Thermally assisted spin-transfer torque dynamics in energy
space. Phys. Rev. B 88, 104405 (2013)
41. Taniguchi, T., Imamura, H.: Thermally assisted spin transfer torque switching in synthetic
free layers. Phys. Rev. B 83, 054432 (2011)
42. Thomas, L., Jan, G., Le, S., Wang, P.-K.: Quantifying data retention of perpendicular spin-
transfer-torque magnetic random access memory chips using an effective thermal stability
factor method. Appl. Phys. Lett. 106, 162402 (2015)
43. Hirayama, E., Sato, H., Kanai, S., Matsukura, F., Ohno, H.: Magnetic field and current
magnetization reversal in elliptic CoFeB/MgO magnetic tunnel junctions with perpendicular
magnetic easy axis. IEEE Magn. Lett. 7, 3104004 (2016)
44. Sun, J.Z., Robertazzi, R.P., Nowak, J., Trouilloud, P.L., Hu, G., Abraham, D.W., Gaidis,
M.C., Brown, S.L., O’Sullivan, E.J., Gallagher, W.J., Worledge, D.C.: Effect of subvolume
excitation and spin-torque efficiency on magnetic switching. Phys. Rev. B 84, 064413 (2011)
45. Sun, J.Z., Trouilloud, P.L., Gajek, M.J., Nowak, J., Robertazzi, R.P., Hu, G., Abraham,
D.W., Gaidis, M.C., Brown, S.L., O’Sullivan, E.J., Gallagher, W.J., Worledge, D.C.: Size
dependence of spin-torque induced magnetic switching in CoFeB-based perpendicular
magnetization tunnel junctions. J. Appl. Phys. 111, 07C711 (2012)
46. Sun, J.Z., Brown, S.L., Chen, W., Delenia, E.A., Gaidis, M.C., Harms, J., Hu, G., Jiang, X.,
Kilaru, R., Kula, W., Lauer, G., Liu, L.Q., Murthy, S., Nowak, J., O’Sullivan, E.J., Parkin,
S.S.P., Robertazzi, R.P., Rice, P.M., Sandhu, G., Topuria, T., Worledge, D.C.: Spin-torque
switching efficiency in CoFeB-MgO based tunnel junctions. Phys. Rev. B 88, 104426 (2013)
47. Apalkov, D.M., Visscher, P.B.: Spin-torque switching: Fokker-Planck rate calculation. Phys.
Rev. B 72, 180405 (2005)
48. Parkin, S.S.P., Kaiser, C., Panchula, A., RiceandBrian Hughes, P.M., Samant, M., Yang, S.-
H.: Giant tunnelling magnetoresistance at room temperature with MgO(100) tunnelbarriers.
Nat. Mater. 3, 862 (2004)
49. Yuasa, S., Nagahama, T., Fukushima, A., Suzukiand Koji Ando, Y.: Giant room-temperature
magnetoresistance in single-crystal Fe/MgO/Fe magnetic tunnel junctions. Nat. Mater. 3, 868
(2004)
50. Butler, W.H., Zhang, X.-G., Schulthess, T.C., MacLaren, J.M.: Spin-dependent tunneling
conductance of Fe|MgO|Fe sandwiches. Phys. Rev. B 63, 054416 (2001)
51. Zhang, X.-G., Butler, W.H.: Large magnetoresistance in bcc Co/MgO/Co and FeCo/M-
gO/FeCo tunnel junctions. Phys. Rev. B 70, 172407 (2004)
52. Ikeda, S., Miura, K., Yamamoto, H., Mizunuma, K., Gan, H.D., Endo, M., Kanai, S.,
Hayakawa, J., Matsukura, F., Ohno, H.: A perpendicular-anisotropy CoFeB-MgO magnetic
tunnel junction. Nat. Mater. 9, 721 (2010)
53. Worledge, D.C., Hu, G., Abraham, D.W., Sun, J.Z., Trouilloud, P.L., Nowak, J., Brown, S.,
Gaidis, M.C., O’Sullivan, E.J., Robertazzi, R.P.: Spin torque switching of perpendicular
Ta|CoFeB|MgO-based magnetic tunnel junctions. Appl. Phys. Lett. 98, 022501 (2011)
54. Ikeda, S., Hayakawa, J., Ashizawa, Y., Lee, Y.M., Miura, K., Hasegawaand, H., Tsunoda,
M., Matsukura, F., Ohno, H.: Tunnel magnetoresistance of 604% at 300 K by supression of
Ta diffusion inCoFeB/MgO/CoFeB pseudo-spin-valves annealed at high temperature. Appl.
Phys. Lett. 93, 082508 (2008)
1358 A. D. Kent et al.

55. Nowak, J.J., Robertazzi, R.P., Sun, J.Z., Hu, G., Park, J.H., Lee, J.H., Annunziata, A.J., Lauer,
G.P., Kothandaraman, C., O’Sullivan, E.J., Trouilloud, P.L., Kim, Y., Worledge, D.C.: Voltage
and size dependence on write-error-rates in STT MRAM down to 11nm junction size. IEEE
Magn. Lett. 7, 3102604 (2016)
56. Watanabe, K., Jinnai, B., Fukami, S., Sato, H., Ohno, H.: Shape anisotropy revisited in single-
digit nanometer magnetic tunnel junctions. Nat. Commun. 9, 663 (2018)
57. Hayakawa, J., Ikeda, S., Lee, Y.M., Sasaki, R., Meguro, T., Matsukura, F., Takahashi,
H., Ohno, H.: Current-driven magnetization switching in CoFeB/MgO/CoFeB magnetic
tunneljunctions. Jpn. J. Appl. Phys. 44, 1267 (2005)
58. Sun, J.Z., Chen, L., Suzuki, Y., Parkin, S.S.P., Koch, R.H.: Thermally activated sweep-rate
dependence of magnetic switching field in nanostructured current-perpendicular spin-valves.
J. Magn. Magn. Mater. 247, L237 (2002)
59. Hosomi, M., Yamagishi, H., Yamamoto, T., Bessho, K., Higo, Y., Yamane, K., Yamada, H.,
Shoji, M., Hachino, H., Fukumoto, C., Nagao, H., Kano, H.: A novel nonvolatile memory
with spin torque transfer magnetization switching: Spin-RAM. In: IEEE. IEDM 2005. IEEE
0-7803-9269-8/05 (2005)
60. Koch, R.H., Katine, J.A., Sun, J.Z.: Time-resolved reversal of spin-transfer switching in a
nanomagnet. Phys. Rev. Lett. 92, 088302 (2004)
61. Sun, J.Z., Kuan, T.S., Katine, J.A., Koch, R.H.: Spin angular momentum transfer in a current-
perpendicular spin-valve nanomagnet. Proc. SPIE 5359, 445 (2004)
62. Tulapurkar, A.A., Devolder, T., Yagami, K., Crozat, P., Chappert, C., Fukushima, A.,
Suzuki, Y.: Subnanosecond magnetization reversal in magnetic nanopillars by spin angular
momentum transfer. Appl. Phys. Lett. 85, 5358 (2005)
63. Bedau, D., Liu, H., Bouzaglou, J.J., Sun, J.Z., Katine, J.A., Mangin, S., Kent, A.D.: Ultrafast
spin-transfer switching in spin valve nanopillars with perpendicular anisotropy. Appl. Phys.
Lett. 96(2), 022514 (2010)
64. Bedau, D.B., Liu, H., Sun, J.Z., Katine, J.A., Mangin, S., Kent, A.D.: Spin-transfer pulse
switching: from the dynamic to the thermally activated regime. Appl. Phys. Lett. 97(26),
262502 (2010)
65. Sun, J.Z., Gaidis, M.C., O’Sullivan, E.J., Joseph, E.A., Hu, G., Abraham, D.W., Nowak, J.J.,
Trouilloud, P.L., Lu, Y., Brown, S.L., Worledge, D.C., Gallagher, W.J.: A three-terminal
spin-torque-driven magnetic switch. Appl. Phys. Lett. 95, 083506 (2009)
66. Zhang, C., Fukami, S., Sato, H., Matsukura, F., Ohno, H.: Spin-orbit torque induced
magnetization switching in nano-scale Ta/CoFeB/MgO. Appl. Phys. Lett. 107, 012401 (2015)
67. Fukami, S., Anekawa, T., Ohkawara, A., Zhang, C., Ohno, H.: A sub-ns three-terminal spin-
orbit torque induced switching device. 2016 Symposium on VLSI Technology, Paper 16-5
(2016)
68. Chaves-O’Flynn, G.D., Vanden-Eijnden, E., Stein, D.L., Kent, A.D.: Thermal stability of
the magnetization in perpendicularly magnetized thin film nanomagnets. J. Appl. Phys. 113,
023912 (2013)
69. Chaves-O’Flynn, G.D., Wolf, G., Sun, J.Z., Kent, A.D.: Thermal stability of magnetic states
in circular thin-film nanomagnets with large perpendicular magnetic anisotropy. Phys. Rev.
Appl. 4, 024010 (2015)
70. Sun, J.Z.: Spin-transfer torque switched magnetic tunnel junctions in magnetic random access
memory. SPIE Conf. 9931, 37 (2016)
71. Nowak, J.J., Robertazzi, R.P., Sun, J.Z., Hu, G., Abraham, D.W., Trouilloud, P.L., Brown, S.,
Gaidis, M.C., O’Sullivan, E.J., Gallagher, W.J., Worledge, D.C.: Demonstration of ultralow
bit error rates for spin-torque magnetic random-access memory with perpendicular magnetic
anisotropy. IEEE Magn. Lett. 2, 3000204 (2011)
72. Kent, A.D., Worledge, D.C.: A new spin on magnetic memories. Nat. Nanotechnol. 10(3),
187–191 (2015)
73. Kent, A.D., Özyilmaz, B., del Barco, E.: Spin-transfer-induced precessional magnetization
reversal. Appl. Phys. Lett. 84, 3897 (2004)
27 Magnetization Dynamics 1359

74. Liu, L., Lee, O.J., Gudmundsen, T.J., Ralph, D.C., Buhrman, R.A.: Current-induced switching
of perpendicularly magnetized magnetic layers using spin torque from the spin Hall effect.
Phys. Rev. Lett. 109, 096602 (2012)
75. Avci, C.O., Garello, K., Nistor, C., Godey, S., Ballesteros, B., Mugarza, A., Barla, A.,
Valvidares, M., Pellegrin, E., Ghosh, A., Miron, I.M., Boulle, O., Auffret, S., Gaudin, G.,
Gambardella, P.: Field-like and antidamping spin-orbit torques in as-grown and annealed
Ta/CoFeB/MgO layers. Phys. Rev. B 89, 214419 (2014)
76. Miron, I.M., Gaudin, G., Auffret, S., Rodmacq, B., Schuhl, A., Pizzini, S., Vogel, J.,
Gambardella, P.: Current-driven spin torque induced by the Rashba effect in a ferromagnetic
metal layer. Nat. Mater. 9, 230 (2010)
77. Avci, C.O., Garello, K., Miron, I.M., Gaudin, G., Auffret, S., Boulle, O., Gambardella, P.:
Magnetization switching of an MgO/Co/Pt layer by in-plane current injection. Appl. Phys.
Lett. 100, 212404 (2012)
78. Lee, O.J., Liu, L.Q., Pai, C.F., Li, Y., Tseng, H.W., Gowtham, P.G., Park, J.P., Ralph, D.C.,
Buhrman, R.A.: Central role of domain wall depinning for perpendicular magnetization
switching driven by spin torque from the spin Hall effect. Phys. Rev. B 89, 024418 (2014)
79. Garello, K., Avci, C.O., Miron, I.M., Baumgartner, M., Ghosh, A., Auffret, S., Boulle, O.,
Gaudin, G., Gambardella, P.: Ultrafast magnetization switching by spin-orbit torques. Appl.
Phys. Lett. 105, 212402 (2014)
80. Cubukcu, M., Boulle, O., Drouard, M., Garello, K., Avci, C.O., Miron, I.M., Langer, J.,
Ocker, B., Gambardella, P.: Spin-orbit torque magnetization switching of a three-terminal
perpendicular magnetic tunnel junction. Appl. Phys. Lett. 104, 042406 (2014)
81. Pinna, D., Stein, D.L., Kent, A.D.: Spin-torque oscillators with thermal noise: a constant
energy orbit approach. Phys. Rev. B 90, 174405 (2014)
82. Pinna, D., Kent, A.D., Stein, D.L.: Thermally assisted spin-transfer torque dynamics in energy
space. Phys. Rev. B 88, 104405 (2013)
83. Liu, H., Bedau, D., Backes, D., Katine, J.A., Langer, J., Kent, A.D.: Ultrafast switching in
magnetic tunnel junction based orthogonal spin transfer devices. Appl. Phys. Lett. 97(24),
242510 (2010)
84. Rowlands, G.E., Rahman, T., Katine, J.A., Langer, J., Lyle, A., Zhao, H., Alzate, J.G.,
Kovalev, A.A., Tserkovnyak, Y., Zeng, Z.M., Jiang, H.W., Galatsis, K., Huai, Y.M., Khalili
Amiri, P., Wang, K.L., Krivorotov, I.N., Wang, J.-P.: Deep subnanosecond spin torque
switching in magnetic tunnel junctions with combined in-plane and perpendicular polarizers.
Appl. Phys. Lett. 98(10), 102509 (2011)
85. Lee, O.J., Pribiag, V.S., Braganca, P.M., Gowtham, P.G., Ralph, D.C., Buhrman, R.A.:
Ultrafast switching of a nanomagnet by a combined out-of-plane and in-plane polarized spin
current pulse. Appl. Phys. Lett. 95(1), 012506 (2009)
86. Beaujour, J.M.L., Bedau, D.B., Liu, H., Rogosky, M.R., Kent, A.D.: Spin-transfer in
nanopillars with a perpendicularly magnetized spin polarizer. Proc. SPIE – Int. Soc. Opt.
Eng. 7398, 73980D (2009)
87. Papusoi, C., Delat́, B., Rodmacq, B., Houssameddine, D., Michel, J.-P., Ebels, U., Sousa,
R.C., Buda-Prejbeanu, L., Dieny, B.: 100 ps precessional spin-transfer switching of a planar
magnetic random access memory cell with perpendicular spin polarizer. Appl. Phys. Lett. 95,
072506 (2009)
88. Park, J., Ralph, D.C., Buhrman, R.A.: Fast deterministic switching in orthogonal spin torque
devices via the control of the relative spin polarizations. Appl. Phys. Lett. 103(25), 252406
(2013)
89. Rowlands, G.E., Ryan, C.A., Ye, L., Rehm, L., Pinna, D., Kent, A.D., Ohki, T.A.: A cryogenic
spin-torque memory element with precessional magnetization dynamics. Sci. Rep. 9, 803
(2019)
90. Pinna, D., Ryan, C.A., Ohki, T., Kent, A.D.: Reliable spin-transfer torque driven precessional
magnetization reversal with an adiabatically decaying pulse. Phys. Rev. B 93, 184412
(2016)
1360 A. D. Kent et al.

91. Houssameddine, D., Ebels, U., Dela et, B., Rodmacq, B., Firastrau, I., Ponthenier, F., Brunet,
M., Thirion, C., Michel, J.-P., Prejbeanu buda, L., Cyrille, M.C., Redon, O., Dieny, B.: Spin-
torque oscillator using a perpendicular polarizer and a planar free layer. Nat. Mater. 6, 447
(2007)
92. Tsoi, M., Jansen, A.G.M., Bass, J., Chiang, W.-C., Seck, M., Tsoi, V., Wyder, P.: Excitation
of a magnetic multilayer by an electric current. Phys. Rev. Lett. 80, 4281 (1998)
93. Rippard, W.H., Pufall, M.R., Silva, T.J.: Quantitative studies of spin-momentum-transfer-
induced excitation in Co/Cu multilayer films using point-contact spectroscopy. Appl. Phys.
Lett. 82, 1260 (2003)
94. Rippard, W.H., Pufall, M.R., Kaka, S., Russek, S.E., Silva, T.J.: Direct-current induced
dynamics in Co90 Fe10 /Ni80 Fe20 pointcontacts. Phys. Rev. Lett. 92, 027201 (2004)
95. Pufall, M.R., Rippard, W.H., Kaka, S., Silva, T.J., Russek, S.E.: Frequency modulation of
spin-transfer oscillators. Appl. Phys. Lett. 86, 082506 (2005)
96. Rippard, W.H., Pufall, M.R., Kaka, S., Silva, T.J., Russek, S.E., Katine, J.A.: Injection locking
and phase control of spin transfer nano-oscillators. Phys. Rev. Lett. 95, 067203 (2005)
97. Slonczewski, J.C.: Excitation of spin waves by an electric current. J. Magn. Magn. Mater.
195(2), L261–L268 (1999)
98. Slavin, A., Tiberkevich, V.: Spin wave mode excited by spin-polarized current in a magnetic
nanocontact is a standing self-localized wave bullet. Phys. Rev. Lett. 95(23), 237201 (2005)
99. Hoefer, M.A., Ablowitz, M.J., Ilan, B., Pufall, M.R., Silva, T.J.: Theory of magnetodynamics
induced by spin torque in perpendicularly magnetized thin films. Phys. Rev. Lett. 95, 267206
(2005)
100. Hoefer, M., Silva, T., Stiles, M.: Model for a collimated spin-wave beam generated by a
single-layer spin torque nanocontact. Phys. Rev. B 77(14), 144401 (2008)
101. Bonetti, S., Tiberkevich, V., Consolo, G., Finocchio, G. Muduli, P., Mancoff, F., Slavin, A.,
Åkerman, J.: Experimental evidence of self-localized and propagating spin wave modes in
obliquely magnetized current-driven nanocontacts. Phys. Rev. Lett. 105, 217204 (2010)
102. Demidov, V.E., Urazhdin, S., Demokritov, S.O.: Direct observation and mapping of spin
waves emitted by spin-torque nano-oscillators. Nat. Mater. 9, 984–988 (2010)
103. Madami, M., Bonetti, S., Consolo, G., Tacchi, G., Carlotti, S., Gubbiotti, G., Mancoff, F.B.,
Yar, M.A., Åkerman, J.: Direct observation of a propagating spin wave induced by spin-
transfer torque. Nat. Nanotechnol. 6, 635 (2011)
104. Madami, M., Iacocca, E., Sani, S., Gubbiotti, G., Tacchi, S., Dumas, R.K., Åkerman, J.,
Carlotti, G.: Propagating spin waves excited by spin-transfer torque: A combined electrical
and optical study. Phys. Rev. B 92, 024403 (2015)
105. Backes, D., Macià, F., Bonetti, S., Kukreja, R., Ohldag, H., Kent, A.D.: Direct observation
of a localized magnetic soliton in a spin-transfer nanocontact. Phys. Rev. Lett. 115, 127205
(2015)
106. Bonetti, S., Kukreja, R., Chen, Z., Macià, F., Hernàndez, J.M., Eklund, A., Backes, D., Frisch,
J., Katine, J., Malm, G., Urazhdin, S., Kent, A.D., Stöhr, J., Ohldag, H., Dürr, H.A.: Direct
observation and imaging of a spin-wave soliton with p−like symmetry. Nat. Commun. 6,
8889 (2015)
107. Hoefer, M.A., Silva, T.J., Keller, M.W.: Theory for a dissipative droplet soliton excited by a
spin torque nanocontact. Phys. Rev. B 82(5), 054432 (2010)
108. Ivanov, A., Kosevich, A.M.: Bound states of a large number of magnons in a ferromagnet
with single ion anisotropy. Zh. Eksp. Teor. Fiz. 72, 2000 (1977)
109. Kosevich, A.M., Ivanov, B.A., Kovalev, A.S.: Magnetic solitons. Phys. Rep. 194, 117 (1990)
110. Mohseni, S.M., Sani, S.R., Persson, J., Anh Nguyen, T.N., Chung, S., Pogoryelov, Y., Muduli,
P.K., Iacocca, E., Eklund, A., Dumas, R.K., Bonetti, S., Deac, A., Hoefer, M.A., Åkerman, J.:
Spin torque generated magnetic droplet solitons. Science 339, 1295 (2013)
111. Macià, F., Backes, D., Kent, A.D.: Stable magnetic droplet solitons in spin-transfer
nanocontacts. Nat. Nanotechnol. 9(12), 992–996 (2014)
112. Chung, S., Eklund, A., Iacocca, E., Mohseni, S.M., Sani, S.R., Bookman, L., Hoefer, M.A.,
Dumas, R.K., Åkkerman, J.: Magnetic droplet nucleation boundary in orthogonal spin-torque
nano-oscillators. Nat. Commun. 7(3), 11209 (2016)
27 Magnetization Dynamics 1361

113. Zhou, Y., Iacocca, E., Awad, A.A., Dumas, R.K., Zhang, F.C., Åkkerman, J.: Dynamically
stabilized magnetic skyrmions. Nat. Commun. 6, 8193 (2015)
114. Statuto, N., Hernàndez, J.M., Kent, A.D., Macià, F.: Generation and stability of dynamical
skyrmions and droplet solitons. Nanotechnology 29(32), 325302 (2018)
115. Houshang, A., Iacocca, E., Durrenfeld, P., Sani, S.R., Åkkerman, J., Dumas, R.K.: Spin-
wave-beam driven synchronization of nanocontact spin-torque oscillators. Nat. Nanotechnol.
11(3), 280+ (2016)
116. Demidov, V.E., Urazhdin, S., Edwards, E.R.J., Stiles, M.D., McMichael, R.D., Demokritov,
S.O.: Control of magnetic fluctuations by spin current. Phys. Rev. Lett. 107, 107204 (2011)
117. Demidov, V.E., Urazhdin, S., Ulrichs, H., Tiberkevich, V., Slavin, A., Baither, D., Schmitz,
G., Demokritov, S.O.: Magnetic nano-oscillator driven by pure spin current. Nat. Mater.
11(12), 1028–31 (2012)
118. Liu, L., Pai, C.-F., Ralph, D.C., Buhrman, R.A.: Magnetic oscillations driven by the spin hall
effect in 3-terminal magnetic tunnel junction devices. Phys. Rev. Lett. 109, 186602 (2012)
119. Demidov, V.E., Urazhdin, S., Zholud, A., Sadovnikov, A.V., Demokritov, S.O.:
Nanoconstriction-based spin-hall nano-oscillator. Appl. Phys. Lett. 105(17), 172410 (2014)
120. Duan, Z., Smith, A., Yang, L., Youngblood, B., Lindner, J., Demidov, V.E., Demokritov, S.O.,
Krivorotov, I.N.: Nanowire spin torque oscillator driven by spin orbit torques. Nat. Commun.
5, 5616 (2014)
121. Ranjbar, M., Durrenfeld, P., Haidar, M., Iacocca, E., Balinskiy, M., Le, T.Q., Fazlali, M.,
Houshang, A., Awad, A., Dumas, R.K., Åkkerman, J.: Cofeb based spin hall nano-oscillators.
IEEE Magn. Lett. 5, 3000504 (2014)
122. Awad, A.A., Dürrenfeld, P., Houshang, A., Dvornik, M., Iacocca, E., Dumas, R.K., Åkker-
man, J.: Long-rang mutual synchronization of spin hall nano-oscillators. Nat. Phys. 13,
292–299 (2016)
123. Kajiwara, Y., Harii, K., Takahashi, S., Ohe, J., Uchida, K., Mizuguchi, M., Umezawa, H.,
Kawai, H., Ando, K., Takanashi, K., Maekawa, S., Saitoh, E.: Transmission of electrical
signals by spin-wave interconversion in a magnetic insulator. Nature 464, 262–266 (2010)
124. Hamadeh, A., d’Allivy Kelly, O., Hahn, C., Meley, H., Bernard, R., Molpeceres, A.H.,
Naletov, V.V., Viret, M., Anane, A., Cros, V., Demokritov, S.O., Prieto, J.L., Muñoz, M.,
de Loubens, G., Klein, O.: Full control of the spin-wave damping in a magnetic insulator
using spin-orbit torque. Phys. Rev. Lett. 113, 197203 (2014)
125. Collet, M., de Milly, X., d’Allivy Kelly, O., Naletov, V.V., Bernard, R., Bortolotti, P.,
Ben Youssef, J., Demidov, V.E., Demokritov, S.O., Prieto, J.L., Munoz, M., Cros, V., Anane,
A., de Loubens, G., Klein, O.: Generation of coherent spin-wave modes in yttrium iron garnet
microdiscs by spin-orbit torque. Nat. Commun. 7(10377) (2016)
126. Cowburn, R.P., Koltsov, D.K., Adeyeye, A.O., Welland, M.E., Tricker, D.M.: Single-domain
circular nanomagnets. Phys. Rev. Lett. 83, 1042–1045 (1999)
127. Manfrini, M., Kim, J.-V., Petit-Watelot, S., Van Roy, W., Lagae, L., Chappert, C., Devolder, T.:
Propagation of magnetic vortices using nanocontacts as tunable attractors. Nat. Nanotechnol.
9(2), 121–125 (2014)
128. Shinjo, T., Okuno, T., Hassdorf, R., Shigeto, K., Ono, T.: Magnetic vortex core observation
in circular dots of permalloy. Science 289(5481), 930–932 (2000)
129. Thiele, A.A.: Steady-state motion of magnetic domains. Phys. Rev. Lett. 30, 230 (1973)
130. Pribiag, V.S., Krivorotov, I.N., Fuchs, G.D., Braganca, P.M., Ozatay, O., Sankey, J.C., Ralph,
D.C., Buhrman, R.A.: Magnetic vortex oscillator driven by d.c. spin-polarized current. Nat.
Phys. 3(7), 498–503 (2007)
131. Dussaux, A., Georges, B., Grollier, J., Cros, V., Khvalkovskiy, A.V., Fukushima, A., Konoto,
M., Kubota, H., Yakushiji, K., Yuasa, S., Zvezdin, K.A., Ando, K., Fert, A.: Large microwave
generation from current-driven magnetic vortex oscillators in magnetic tunnel junctions. Nat.
Commun. 1(8) (2010)
132. Petit-Watelot, S., Kim, J.-V., Ruotolo, A., Belanovsky, A.D., Skirdkov, P.N., Otxoa, R.M.,
Bouzehouane, K., Grollier, J., Vansteenkiste, A., Van de Wiele, B., Cros, V., Devolder,
T.: Commensurability and chaos in magnetic vortex oscillations. Nat. Phys. 8(9):682–687
(2012)
1362 A. D. Kent et al.

133. Sluka, V., K-kay, A., Deac, A.M., Bürgler, D.E., Schneider, C.M., Hertel, R.: Spin-torque-
induced dynamics at fine-split frequencies in nano-oscillators with two stacked vortices. Nat.
Commun. 6, 6409 (2015)
134. Locatelli, N., Hamadeh, A., Abreu, A.F., Belanovsky, A.D., Skirdkov, P.N., Belanovsky,
A.D., Lebrun, R., Naletov, V.V., Zvezdin, K.A., Munoz, M., Grollier, J., Klein, O., Cros,
V., de Loubens, G.: Efficient synchronization of dipolarly coupled vortex-based spin transfer
nano-oscillators. Sci. Rep. 5(17039) (2015)
135. Rippard, W.H., Pufall, M.R., Kaka, S., Silva, T.J., Russek, S.E., Katine, J.A.: Injection locking
and phase control of spin transfer nano-oscillators. Phys. Rev. Lett. 95, 067203 (2005)
136. Bürgler, D.E., Sluka, V., Lehndorff, R., Deac, A.M., K-kay, A., Hertel, R., Schneider, C.M.:
Injection locking of single-vortex and double-vortex spin-torque oscillators. Proc. SPIE 8100,
810018 (2011)
137. Park, J.P., Eames, P., Engebretson, D.M., Berezovsky, J., Crowell, P.A.: Imaging of spin
dynamics in closure domain and vortex structures. Phys. Rev. B 67,020403 (2003)
138. Choe, S.-B., Acremann, Y., Scholl, A., Bauer, A., Doran, A., Stöhr, J., Padmore, H.A.: Vortex
core-driven magnetization dynamics. Science 304(5669), 420–422 (2004)
139. Stöhr, J., Siegmann, H.C.: Magnetism: From Fundamentals to Nanoscale Dynamics. Springer,
Berlin/Heidelberg (2006)
140. Bigot, J.-Y., Hübner, W., Rasing, T., Chantrell, R. (eds.): Proceedings of the International
Conference UMC 2013. Springer (2013)
141. Tudosa, I., Stamm, C., Kashuba, A.B., King, F., Siegmann, H.C., Stöhr, J., Ju, G., Lu, B.,
Weeler, D.: The ultimate speed of magnetic switching in granular recording media. Nature
428(6985), 831–833 (2004)
142. Hansteen, F., Kimel, A., Kirilyuk, A., Rasing, T.: Nonthermal ultrafast optical control of the
magnetization in garnet films. Phys. Rev. B 73(1), 1–14 (2006)
143. Hansteen, F., Kimel, A., Kirilyuk, A., Rasing, T.: Femtosecond photomagnetic switching of
spins in ferrimagnetic garnet films. Phys. Rev. Lett. 95(4), 1–4 (2005)
144. Kimel, A.V., Kirilyuk, A., Usachev, P.A., Pisarev, R.V., Balbashov, A.M., Rasing, T.: Ultrafast
non-thermal control of magnetization by instantaneous photomagnetic pulses. Nature
435(7042), 655–657 (2005)
145. Stanciu, C.D., Hansteen, F., Kimel, A.V., Kirilyuk, A., Tsukamoto, A., Itoh, A., Rasing, T.:
All-optical magnetic recording with circularly polarized light. Phys. Rev. Lett. 99, 047601
(2007)
146. Lambert, C.-H., Mangin, S., Varaprasad, B.S.D.C.S., Takahashi, Y.K., Hehn, M., Cinchetti,
M., Malinowski, G., Hono, K., Fainman, Y., Aeschlimann, M., Fullerton, E.E.: All-optical
control of ferromagnetic thin films and nanostructures. Science 345(6202), 1337–1340 (2014)
147. Zabel, H., Farle, M.: Magnetic Nanostructures. Springer Tracts in Modern Physics. Springer
(2013). https://link.springer.com/book/10.1007/978-3-642-32042-2
148. Wessels, P., Vogel, A., Tödt, J.-N., Wieland, M., Meier, G., Drescher, M.: Direct observa-
tion of isolated damon-eshbach and backward volume spin-wave packets in ferromagnetic
microstripes. Sci. Rep. 6, 22117 EP (2016)
149. Mesler, B.L., Fischer, P., Chao, W., Anderson, E.H., Kim, D.-H.: Soft x-ray imaging of spin
dynamics at high spatial and temporal resolution. J. Vacuum Sci. Technol. B 25(6), 2598–
2602 (2007)
150. Fischer, P., Ohldag, H.: X-rays and magnetism. Rep. Progress Phys. 78(9), 094501 (2015)
151. Acremann, Y., Strachan, J.P., Chembrolu, V., Andrews, S.D., Tyliszczak, T., Katine, J.A.,
Carey, M.J., Clemens, B.M., Siegmann, H.C., Stohr, J.: Time-resolved imaging of spin
transfer switching: beyond the macrospin concept. Phys. Rev. Lett. 96(21), 217202 (2006)
152. Acremann, Y., Chembrolu, V., Strachan, J.P., Tyliszczak, T., Stöhr, J.: Software defined
photon counting system for time resolved x-ray experiments. Rev. Sci. Instrum. 78(1), 014702
(2007)
153. Bonetti, S., Kukreja, R., Chen, Z., Spoddig, D., Ollefs, K., Schöppner, C., Meckenstock, R.,
Ney, A., Pinto, J., Houanche, R., Frisch, J., Stöhr, J., Dürr, H.A., Ohldag, H.: Microwave soft
x-ray microscopy for nanoscale magnetization dynamics in the 5–10 ghz frequency range.
Rev. Sci. Instrum. 86(9), 093703 (2015)
27 Magnetization Dynamics 1363

154. Acremann, Y., Back, C.H., Buess, M., Portmann, O., Vaterlaus, A., Pescia, D., Melchior,
H.: Imaging precessional motion of the magnetization vector. Science 290(5491), 492–495
(2000)
155. Open, H.P., Hopster, H.: Magnetic Microscopy of Nanostructures. Nanoscience and
Technology. Springer (2005)
156. Beaurepaire, E., Merle, J.C., Daunois, A., Bigot, J.-Y.: Ultrafast spin dynamics in ferromag-
netic nickel. Phys. Rev. Lett. 76, 4250 (1996)
157. Ju, G., Nurmikko, A.V., Farrow, R.F.C., Marks, R.F., Carey, M.J., Gurney, B.A.: Ultrafast time
resolved photoinduced magnetization rotation in a ferromagnetic/antiferromagnetic exchange
coupled system. Phys. Rev. Lett. 82, 3705 (1999)
158. Ju, G., Chen, L., Nurmikko, A.V., Farrow, R.F.C., Marks, R.F., Carey, M.J., Gurney, B.A.:
Coherent magnetization rotation induced by optical modulation in ferromagnetic/antiferro-
magnetic exchange-coupled bilayers. Phys. Rev. B 62, 1171 (2000)
159. van Kampen, M., Jozsa, C., Kohlhepp, J.T., LeClair, P., Lagae, L., de Jonge, W.J.M.,
Koopmans, B.: All-optical probe of coherent spin waves. Phys. Rev. Lett. 88, 227201
(2002)
160. Thiele, J.U., Buess, M., Back, C.H.: Spin dynamics of the antiferromagnetic-to-ferromagnetic
phase transition in ferh on a sub-picosecond time scale. Appl. Phys. Lett. 85, 2857 (2004)
161. Ju, G.P., Hohlfeld, J., Bergman, B., van de Veerdonk, R.J.M., Mryasov, O.N., Kim, J.Y., Wu,
X.W., Weller, D., Koopmans, B.: Ultrafast generation of ferromagnetic order via a laser-
induced phase transformation in ferh thin films. Phys. Rev. Lett. 93, 197403 (2004)
162. Mangin, S., Gottwald, M., Lambert, C.-H., Steil, D., Uhlíř, V., Pang, L., Hehn, M., Alebrand,
S., Cinchetti, M., Malinowski, G., Fainman, Y., Aeschlimann, M., Fullerton, E.E.: Engineered
materials for all-optical helicity-dependent magnetic switching. Nat. Mater 13(3), 286–292
(2014)
163. Lambert, C.-H., Mangin, S., Varaprasad, B.S.D.C.S., Takahashi, Y.K., Hehn, M., Cinchetti,
M., Malinowski, G., Hono, K., Fainman, Y., Aeschlimann, M., Fullerton, E.E.: All-optical
control of ferromagnetic thin films and nanostructures. Science 345(6202), 1337–1340
(2014)
164. Stanciu, C., Hansteen, F., Kimel, A., Kirilyuk, A., Tsukamoto, A., Itoh, A., Rasing, T.: All-
optical magnetic recording with circularly polarized light. Phys. Rev. Lett. 99, 047601 (2007)
165. Battiato, M., Carva, K., Oppeneer, P.M.: Superdiffusive spin transport as a mechanism of
ultrafast demagnetization. Phys. Rev. Lett. 105, 027203 (2010)
166. Melnikov, A., Razdolski, I., Wehling, T.O., Papaioannou, E.T., Roddatis, V., Fumagalli, P.,
Aktsipetrov, O., Lichtenstein, A.I., Bovensiepen, U.: Ultrafast transport of laser-excited spin-
polarized carriers in Au/Fe/MgO(001). Phys. Rev. Lett. 107, 076601 (2011)
167. Kaltenborn, S., Zhu, Y.-H., Schneider, H.C.: Wave-diffusion theory of spin transport in metals
after ultrashort-pulse excitation. Phys. Rev. B 85, 235101 (2012)
168. Alekhin, A., Bürstel, D., Melnikov, A., Diesing, D., Bovensiepen, U.: Ultrafast laser-excited
spin transport in au/fe/mgo (001): relevance of the fe layer thickness. In: Bigot et al. (2013),
p 214
169. Melnikov, A., Alekhin, A., Bürstel D., Diesing, D., Wehling, T.O., Rungger, I., Stamenova,
M., Sanvito, S., Bovensiepen, U.: Ultrafast non-local spin dynamics in metallic bi-layers by
linear and non-linear magneto-optics. In: Bigot et al. (2013), p 34
170. Schellekens, A.J., Verhoeven, W., Vader, T.N., Koopmans, B.: Investigating the contribution
of superdiffusive transport to ultrafast demagnetization of ferromagnetic thin films. Appl.
Phys. Lett. 102(25), 252408 (2013)
171. Ando, K., Takahashi, S., Ieda, J., Kajiwara, Y., Nakayama, H., Yoshino, T., Harii, K.,
Fujikawa, Y., Matsuo, M., Maekawa, S., Saitoh, E.: Inverse spin-hall effect induced by spin
pumping in metallic system. J. Appl. Phys. 109(10), 103193 (2011)
172. Hoffmann, A.: Spin hall effects in metals. IEEE Trans. Magn. 49, 5172 (2013)
173. Kampfrath, T., Battiato, M., Maldonado, P., Eilers, G., Notzold, J., Mahrlein, S., Zbarsky, V.,
Freimuth, F., Mokrousov, Y., Blügel, S., Wolf, M., Radu, I., Oppeneer, P.M., Münzenberg,
M.: Terahertz spin current pulses controlled by magnetic heterostructures. Nat. Nano 8(4),
256–260 (2013)
1364 A. D. Kent et al.

174. Seifert, T., Jaiswal, S., Martens, U., Hannegan, J., Braun, L., Maldonado, P., Freimuth,
F., Kronenberg, A., Henrizi, J., Radu, I., Beaurepaire, E., Mokrousov, Y., Oppeneer, P.M.,
Jourdan, M., Jakob, G., Turchinovich, D., Hayden, L.M., Wolf, M., Münzenberg, M., Kläui,
M., Kampfrath, T.: Efficient metallic spintronic emitters of ultrabroadband terahertz radiation.
Nat. Photon 10(7), 483–488 (2016)
175. Malinowski, G., Dalla Longa, F., Rietjens, J.H.H., Paluskar, P.V., Huijink, R., Swagten,
H.J.M., Koopmans, B.: Control of speed and efficiency of ultrafast demagnetization by direct
transfer of spin angular momentum. Nat. Phys. 4(11), 855–858 (2008)
176. Rudolf, D., La-O-Vorakiat, C., Battiato, M., Adam, R., Shaw, J.M., Turgut, E., Maldonado,
P., Mathias, S., Grychtol, P., Nembach, H.T., Silva, T.J., Aeschlimann, M., Kapteyn, H.C.,
Murnane, M.M., Schneider, C.M., Oppeneer, P.M.: Ultrafast magnetization enhancement in
metallic multilayers driven by superdiffusive spin current. Nat. Commun. 3, 1037 EP (2012)
177. Turgut, E., La-o vorakiat, C., Shaw, J.M., Grychtol, P., Nembach, H.T., Rudolf, D., Adam, R.,
Aeschlimann, M., Schneider, C.M., Silva, T.J., Murnane, M.M., Kapteyn, H.C., Mathias, S.:
Controlling the competition between optically induced ultrafast spin-flip scattering and spin
transport in magnetic multilayers. Phys. Rev. Lett. 110, 197201 (2013)
178. Eschenlohr, A., Battiato, M., Maldonado, P., Pontius, N., Kachel, T., Holldack, K., Mitzner,
R., Föhlisch, A., Oppeneer, P.M., Stamm, C.: Ultrafast spin transport as key to femtosecond
demagnetization. Nat. Mater 12(4), 332–336 (2013)
179. Eschenlohr, A., Battiato, M., Maldonado, P., Pontius, N., Kachel, T., Holldack, K., Mitzner,
R., Fohlisch, A., Oppeneer, P.M., Stamm, C.: Reply to ’optical excitation of thin magnetic
layers in multilayer structures’. Nat. Mater 13(2), 102–103 (2014)
180. Khorsand, A.R., Savoini, M., Kirilyuk, A., Rasing, T.: Optical excitation of thin magnetic
layers in multilayer structures. Nat. Mater 13(2), 101–102 (2014)
181. Choi, G.-M., Min, B.-C., Lee, K.-J., Cahill, D.G.: Spin current generated by thermally driven
ultrafast demagnetization. Nat. Commun. 5, 4334 EP (2014)
182. He, W., Zhu, T., Zhang, X.-Q., Yang, H.-T., Cheng, Z.-H.: Ultrafast demagnetization
enhancement in cofeb/mgo/cofeb magnetic tunneling junction driven by spin tunneling
current. Sci. Rep. 3, 2883 EP (2013)
183. Savoini, M., Piovera, C., Rinaldi, C., Albisetti, E., Petti, D., Khorsand, A.R., Duò, L., Dallera,
C., Cantoni, M., Bertacco, R., Finazzi, M., Carpene, E., Kimel, A.V., Kirilyuk, A., Rasing,
T.: Bias-controlled ultrafast demagnetization in magnetic tunnel junctions. Phys. Rev. B 89,
140402 (2014)
184. Schellekens, A.J., Kuiper, K.C., de Wit, R.R.J.C., Koopmans, B.: Ultrafast spin-transfer
torque driven by femtosecond pulsed-laser excitation. Nat. Commun. 5, 4333 EP (2014)
185. Radu, I., Vahaplar, K., Stamm, C., Kachel, T., Pontius, N., Durr, H.A., Ostler, T.A., Barker,
J., Evans, R.F.L., Chantrell, R.W., Tsukamoto, A., Itoh, A., Kirilyuk, A., Rasing, T., Kimel,
A.V.: Transient ferromagnetic-like state mediating ultrafast reversal of antiferromagnetically
coupled spins. Nature 472(7342), 205–208 (2011)
186. Graves, C.E., Reid, A.H., Wang, T., Wu, B., de Jong, S., Vahaplar, K., Radu, I., Bernstein,
D.P., Messerschmidt, M., Müller, L., Coffee, R., Bionta, M., Epp, S.W., Hartmann, R.,
Kimmel, N., Hauser, G., Hartmann, A., Holl, P., Gorke, H., Mentink, J.H., Tsukamoto,
A., Fognini, A., Turner, J.J., Schlotter, W.F., Rolles, D., Soltau, H., Strüder, L., Acremann,
Y., Kimel, A.V., Kirilyuk, A., Rasing, T., Stöhr, J., Scherz, A.O., Dürr, H.A.: Nanoscale
spin reversal by non-local angular momentum transfer following ultrafast laser excitation in
ferrimagnetic gdfeco. Nat. Mater 12(4), 293–298 (2013)
187. Kukreja, R., Bonetti, S., Chen, Z., Backes, D., Acremann, Y., Katine, J.A., Kent, A.D., Dürr,
H.A., Ohldag, H., Stöhr, J.: X-ray detection of transient magnetic moments induced by a spin
current in Cu. Phys. Rev. Lett. 115, 096601 (2015)
188. Hellman, F., Hoffmann, A., Tserkovnyak, Y., Beach, G.S.D., Fullerton, E.E., Leighton, C.,
MacDonald, A.H., Ralph, D.C., Arena, D.A., Dürr, H.A., Fischer, P., Grollier, J., Heremans,
J.P., Jungwirth, T., Kimel, A.V., Koopmans, B., Krivorotov, I.N., May, S.J., Petford-Long,
A.K., Rondinelli, J.M., Samarth, N., Schuller, I.K., Slavin, A.N., Stiles, M.D., Tchernyshyov,
O., Thiaville, A., Zink, B.L.: Interface-induced phenomena in magnetism. Rev. Mod. Phys.
89, 025006 (2017)
27 Magnetization Dynamics 1365

Andrew D. Kent is a Professor of Physics and Founding Director


of the Center for Quantum Phenomena at NYU. He received a
B.Sc. with distinction in Applied and Engineering Physics at Cor-
nell University in 1982 and his Ph.D. from Stanford University
in Applied Physics in 1988. His research interests are the physics
of magnetic nanostructures, nanomagnetic devices, and magnetic
information storage.

Hendrik Ohldag is a Staff Scientist at Lawrence Berkeley


National Laboratory, as well as an Adjunct Professor for Physics
at UC Santa Cruz and Materials Science and Engineering at
Stanford University. He received his Doctorate from the Institute
of Applied Physics at the University of Duesseldorf, Germany.
His research focuses on the study of magnetic materials with
X-ray-based spectromicroscopy.

Hermann A. Dürr is a Professor of Instrumentation and Accel-


erator Physics at Uppsala University, Sweden. He received his
Ph.D. from the University of Bayreuth, Germany, in 1989 and
worked at Daresbury Laboratory, UK; BESSY GmbH, Germany;
and SLAC National Accelerator Laboratory, USA. He uses soft
X-rays for ultrafast magnetism and is currently discovering the
joys and pitfalls of structural dynamics with electrons.

Jonathan Z. Sun is a Research Staff Member at the IBM T. J.


Watson Research Center in New York. He received his Ph.D. and
M.S. from Applied Physics at Stanford University in 1989 and
1986 and his B.S. in Physics from Fudan University in Shanghai,
China, in 1984. His research interests include device and materials
physics in magnetic and superconducting systems.
Part IV
Applications
Permanent Magnet Materials and
Applications 28
Karl-Hartmut Müller, Simon Sawatzki , Roland Gauß,
and Oliver Gutfleisch

Contents
Permanent Magnets and Physics Behind Them . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1374
Stray Fields and Demagnetizing Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1376
On the Thermodynamics of Magnets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1379
Hysteresis Curves and Magnetization Processes of Permanent Magnets . . . . . . . . . . . . . . 1384
Permanent Magnet Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1387
Hard Magnetic Steels and Alnico . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1390
Hard Ferrite . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1392
Rare-Earth Magnets Based on SmCo5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1394
Sm2 Co17 -Based Magnets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1395
Nd2 Fe14 B-Based Magnets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1397
Other Permanent Magnet Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1401
Resource Constraints . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1405
Future Permanent Magnet Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1409
Permanent Magnet Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1412
Permanent Magnet Assemblies and Their Dynamic Application with
Mechanical Recoil . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1413
Dynamic Applications with Active Recoil . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1421
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1425

K.-H. Müller
IFW Dresden, Institute for Metallic Materials, Dresden, Germany
S. Sawatzki
Technische Universität Darmstadt, Materialwissenschaft, Darmstadt, Germany
Vacuumschmelze GmbH & Co.KG, Hanau, Germany
e-mail: simon.sawatzki@vacuumschmelze.com
R. Gauß
EIT RawMaterials GmbH, Berlin, Germany
e-mail: roland.gauss@eitrawmaterials.eu
O. Gutfleisch ()
Technische Universität Darmstadt, Materialwissenschaft, Darmstadt, Germany
e-mail: oliver.gutfleisch@tu-darmstadt.de
© Springer Nature Switzerland AG 2021 1369
J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_29
1370 K.-H. Müller et al.

Abstract

A permanent magnet (PM) is able to carry a macroscopic magnetization whose


magnitude and direction are persistent and resist applied magnetic fields opposite
or transverse to the magnetization. This is understood in terms of the rela-
tivistic quantum physics, solid-state physics, and physics of thermodynamically
metastable states. Permanent magnetism is also a field of material science
because resistance to magnetization reversal is governed by details of the
microstructure of the PM material. The mechanisms behind that relationship
are not yet well understood. Today hard ferrites represent the lion’s share by
mass of the PM shipped worldwide, and Nd-Fe-B materials dominate by value
the PM world market. The latter are indispensable for many modern static as
well as dynamic applications such as electric motors in industry and household
applications, hard disk drives, various automotive applications, wind power,
electric bikes, air conditioning, and speakers. The Nd-Fe-B materials are highly
critical in terms of resource criticality, which can be explained by a high risk in
the raw material supply as well as the great economic importance of the magnet
material. Current research tries to reduce the need of heavy rare-earth elements
by advanced preparation routes for Nd-Fe-B magnets. Partial substitution of Nd
by more available elements such as Ce or La and the search for novel materials, in
particular those containing Mn, are expected to fill the gap between the magnetic
properties of Nd-Fe-B and ferrites. Novel PM materials with an energy product
(BH )max greater than that of sintered Nd-Fe-B are not in sight.

List of Symbols and Abbreviations

A Exchange constant
Ag Cross-sectional areas of a cavity
Am Cross-sectional areas of a magnet
AC Alternating current
B Vector of the flux density (or induction)
B(r) Flux density (or induction) depending on the position
r
B Magnitude of the flux density (or induction)
(BH )max “Energy product,” i.e., the maximum value of (−BH )
of a homogeneously magnetized body measured on its
major hysteresis loop, sometimes called energy density
(BH )u “Recoil product,” i.e., the maximum energy product in
a dynamic application with mechanical recoil
C Torque constant of a motor
d-HDDR Dynamic hydrogenation-disproportionation-
desorption-recombination
d Distance of a small air gap
Dc Critical single-domain particle diameter
28 Permanent Magnet Materials and Applications 1371

Ds Domain width on the surface of an easy-axis magnetic


phase
DC Direct current
Eg Magnetostatic energy in the air gap
Em Magnetostatic energy created by a magnetized body
emf Electromagnetic force (voltage)
Eout Magnetostatic energy created by a magnetized body
outside of that body
fr Rotational frequency of the rotor
fc Cyclotron frequency
F Force
Fh Holding force
FL Lorentz force
Fm Force of a magnetic moment in a field gradient

G(p, T , H ) Equilibrium Gibbs free energy depending on pressure

G̃(p, T , H ; λ) Gibbs free energy of a manifold of constrained equi-
librium states where λ is a parameter symbolically
representing these states
 
Ĝ(A, K, H ; [M(r )]) Landau-type potential non-locally depending on M(r)
GB Grain boundary (also used for the intergranular mate-
rial)
GBD Grain boundary diffusion
H The vector of the magnetic field (or magnetic field
strength)
H (r) Magnetic field depending on the position r
H V Volume average of the magnetic field

H Magnitude of the external magnetic field

H External magnetic field
Ha Anisotropy field
Hc Coercivity (or coercive field)
Hg Magnetic field in a certain air gap
Hr Magnetic field of a point or a line dipole along the
dipole axis
Hθ Magnetic field of a point or a line dipole along the
tilting angle θ
Hφ Magnetic field of a point or a line dipole along the
azimuth angle φ
HDDR Hydrogenation-disproportionation-desorption-
recombination
I Current
K1 Second-order anisotropy constant of a magnetically
uniaxial material
K1m Effective anisotropy constant based on shape anisotropy
kB Boltzmann constant
1372 K.-H. Müller et al.

li Intrinsic magnetic lengths (of general type)


lκ Anisotropy-related exchange length
L Length of a segment or a winding
LED Light-emitting diode
M The vector of magnetization
M(r) Magnetization depending on the position r
M(t) Magnitude of the magnetization depending on time t

M(r, T , H ) Magnetization depending on the position r in the body,
temperature T , and the history of the external magnetic

field H applied to that body
m Magnetic moment
me Mass of the electron
Mr Remanence (or remanent magnetization)
Mr0 Ideal remanence of a non-textured material
Ms Magnitude of the spontaneous magnetization of a
phase
Ms The vector of spontaneous magnetization
Msat Saturation magnetization of a material
MRI Magnetic resonance imaging
n Number of segments or rods in a Halbach array or a
magnetic mangle, respectively
np Number of poles
nw Number of windings
N Demagnetization tensor
N Demagnetization factor measured in a given direction
N1 , N2 , N3 The eigenvalues of N
N North pole
Neff Effective demagnetization factor (parameter in a phe-
nomenological formula of Hc )
Ng Scale-independent geometric factor of a magnetic cir-
cuit
NMR Nuclear magnetic resonance
p Pressure
P1 Working point of a permanent magnet
P2 Working point of a permanent magnet after applying a
magnetic field or changing the air gap
P3 Position within a magnet assembly
P Power of a PM machine
Pa Acoustic power
PLP Pressless (sinter) process
PM Permanent magnet
PSA Post-sinter annealing
q Electrical charge of an electron
28 Permanent Magnet Materials and Applications 1373

R Electrical resistance of a coil


r1 , r2 , r3 Inner and outer radius of a magnet assembly as well as
its thickness
rr Radius of the rotor
rw Radius of the rotor windings
REE Rare-earth element
REO Rare-earth oxide
S Magnetic viscosity
S South pole
Sv Viscosity coefficient (or fluctuation field)
t Time
T Temperature
Tc Curie temperature
t0 Time parameter characterizing a magnetic viscosity
experiment
U Voltage
V Volume
va Activation volume
VSM Vibrating sample magnetometer

Greek Symbols (and Abbreviations)

α Parameter in a phenomenological formula of Hc


α Temperature coefficient of Mr
β Temperature coefficient of Hc
δw Bloch wall width
η Tilting angle of the magnetization in a magnet assembly
ε Efficiency of a magnet assembly
γ Line dipole
Γ Torque
κ Magnetic hardness parameter
λ Parameter symbolically representing constrained states in thermodynam-
ics
μ0 Permeability of the vacuum
μr Recoil permeability in a dynamic application with mechanical recoil
ω Angular velocity
φ Azimuth angle
Φ Magnetic flux
ρ Specific electrical resistivity
θ Tilting angle (or angular position)
χ Susceptibility
χirr Irreversible part of the magnetic susceptibility
1374 K.-H. Müller et al.

Permanent Magnets and Physics Behind Them

A permanent magnet (PM), also called hard magnet, is a piece of material that cre-
ates a magnetic field outside that material without the need for any external source of
power. In this section, we will briefly summarize the physics of phenomena typically
found for the PM. More details can be found in Chapters “ATO,” “EXC,” “ANI,”
“MPT,” “MIC,” “DOM,” “ELE,” “MMA,” and “MOX.”
An important class of PMs are bodies that carry a persistent magnetization vector

M(r, T , {H }) depending on the position r in the body, the temperature T , and the
 
history, {H }, of the externally applied magnetic field H . This influence of the

history {H } will be discussed in section “On the Thermodynamics of Magnets”.

The vector M(r, T , {H }) is an ingredient of the continuum theory of mag-
netism called micromagnetism or micromagnetics [1, 2, 3]. For a typical PM
material based on a uniaxial (tetragonal, rhombohedral, or hexagonal) magnetic
phase, the simplest version of micromagnetism contains three material properties:
the magnitude of the spontaneous magnetization, Ms ; the so-called second-order
anisotropy constant K1 > 0; and the exchange constant A. These properties
are not constants but functions of temperature and are assumed to be independent

of H . It is known that Ms is based on microscopic magnetic moments that are
closely related to the orbital and spin momentum of valence electrons and are
strongly coupled by exchange interaction, resulting in cooperative phenomena such
as ferromagnetism (as in Fe or Nd2 Fe14 B) or ferrimagnetism (as in (Ba,Sr)Fe12 O19
or in lodestone composed of Fe3 O4 , which is the main source of rock magnetism)
or canted antiferromagnets which show weak ferromagnetism (as in hematite,
α-Fe2 O3 , that has a very small value of Ms but if present in large amounts
contributes to natural magnetic remanence of rocks and, consequently, plays an
important role in paleomagnetism).
A finite magnetization Ms can only exist at room temperature because the
exchange interaction between the concerned spins is very strong. It arises from the
electrostatic (Coulomb) interaction between electrons under the rules of quantum
mechanics, in particular the Pauli exclusion principle [4]. In micromagnetism,
the exchange interaction between neighboring parts of a material is represented
by the exchange parameter A. The persistence of the magnetization of a PM
material in a particular direction is due to magnetic anisotropy represented by
K1 . If in a uniaxial material K1 > 0, such preferred directions are parallel to
the crystallographic c-axis, which is then called the easy axis of magnetization.
The magnetic anisotropy can be based on different microscopic mechanisms. In
modern PM materials, its dominant mechanism has two components. The first
component is the electron orbitals and with them the orbital magnetic moments,
which are coupled to the crystal axes by electrostatic forces (again governed by
quantum mechanics). This was first described by Hans Bethe who investigated
this effect for certain materials, and he related these forces to crystalline electric
fields (Kristallfeld – in German) [5]. The second component is the spin-orbit
interaction that mediates this anisotropy to the spins, which mainly contributes to
28 Permanent Magnet Materials and Applications 1375

the magnetization Ms . The crystalline electric fields obviously do not have rotational
symmetry, and, consequently, the orbital magnetic moments observed in free atoms
or ions are quenched in the solid (at least partially). Hence, magnetic and electric
phenomena are not only closely connected according to the electrodynamics of
Maxwell or Einstein’s theory of relativity but also in terms of exchange interaction
and magnetic anisotropy, which are magnetic phenomena but result from the
electrostatic Coulomb interaction. A different type of magnetic anisotropy, shape
anisotropy, will be discussed in section “Hard Magnetic Steels and Alnico”.
In ferrite magnets based on BaFe12 O19 and SrFe12 O19 , which represent the lion’s
share by mass of the PM shipped worldwide each year [6], the magnetization is
exclusively a superposition of the spin magnetic moments of Fe3+ ions (from 5 of
their 23 electrons). In Nd2 Fe14 B-based materials, which dominate the PM world
market by value [7], besides the contribution of spin magnetic moments, a few
percent of the magnetization is contributed by Nd 4f orbital magnetic moments,
because they are not quenched due to the strong spin-orbit interaction resisting the
crystalline electric fields in the rare earths [8].
The intrinsic properties Ms , A, and K1 of a material lead to the anisotropy-related
exchange length

A
lκ = (1)
K1

and further intrinsic magnetic lengths li of the system that are generally given by

K1
li = lκ Fi (κ) with κ 2 = (2)
μ0 Ms2

where Fi (κ) is a function of the hardness parameter κ [9], relating the anisotropy
energy to the magnetostatic energy, and μ0 is the permeability of free space.
Micromagnetism is a mesoscopic theory, and limits of its validity arise if it predicts
phenomena on length scales li which are not considerably larger than the lattice
constant of the considered magnetic phase.
Alternatively, a piece of superconducting material can also be magnetized and
can be used as a permanent magnet [10,11]. Much higher fields can be realized using
superconducting permanent magnets compared to that of their ferromagnetic
counterparts. Trapped fields as high as 16 T could be achieved at 24 K in the gap
between two YBa2 Cu3 O7−δ superconducting bulk samples (25 mm in diameter)
which were doped with Zr and Ag and placed inside of a reinforcing stainless steel
tube [12]. An even higher trapped field of 17.6 T at 26 K was reported for a stack
of two Ag-doped GdBa2 Cu3 O7−δ superconducting bulk samples, each 25 mm in
diameter, reinforced with shrink-fit stainless steel [13]. However, superconducting
permanent magnets only exist at cryogenic temperatures, and they will not be further
considered in this chapter.
1376 K.-H. Müller et al.

Stray Fields and Demagnetizing Fields

In an idealized permanent magnet, the magnetization M is uniform, it does not


depend on r, and it is rigid that its magnitude and its orientation are insensitive to
applied magnetic fields. The field generated by such an idealized magnet, H (r), can
be obtained using Maxwell’s equations:

∇ × H (r) = 0 (3)

 
∇ ◦ M(r) + H (r) = 0 (4)

where the spatial dependence of M(r) describes the transition from the uniformly
magnetized body to its nonmagnetic environment. To solve (3) and (4) is straightfor-
ward but laborious [14]. As an example, Fig. 1 schematically shows the field H (r)
and the induction or flux density
 
B(r) = μ0 M(r) + H (r) (5)

of a uniformly magnetized cylinder. In many cases, exact results concerning the


field H (r) generated by M can be obtained without any calculation but only by
using general theorems and symmetry considerations, as will be exemplified in the
following:

Fig. 1 Schematic presentation of the fields H and B created by the magnetization M of a


uniformly magnetized cylinder. The symbols • and ∗ mark positions close to the cylinder’s surface
(see text). (Modified after [9, Fig. 2.7, p. 35])
28 Permanent Magnet Materials and Applications 1377

The volume average of the field within a uniformly magnetized body of arbitrary
shape is given by

H V = −N M (6)

[15, 1] where N is a symmetric tensor termed demagnetization tensor with non-


negative eigenvalues N1 , N2 , and N3 called demagnetization factors that satisfy

N1 + N2 + N3 = 1. (7)

It should be noted that the average field H V is parallel to M only if M is parallel


to one of the principal axes of N. In many cases, the principal axes of N can be
inferred because they coincide with the symmetry axes of the body, for example, the
axis of the circular cylinder of Fig. 1 and two axes perpendicular to it. For cylinders
with a very large length-to-diameter ratio, L/∅ → ∞, the axial demagnetization
factor vanishes, Nz = 0, whereas flat cylinders (L/∅ → 0) have Nz = 1 [1], that is,
axially magnetized long or short cylinders show H V = 0 or H V = −M, respec-
tively. By symmetry arguments, it can be shown that for long cylinders, the field
values at the position ∗ in Fig. 1 are H = −M/2 (despite H V = 0!) and, at the
position •, we have H = M/2. In case of a short cylinder, at ∗ and •, H = −M, and
H = 0, respectively. Corresponding values of B can be easily obtained using (5).
For reasons of symmetry, it can be concluded that for symmetric bodies like
those in Fig. 2 (with the exception of the cylinder), each axis is a principal axis of
N with the eigenvalue N = 1/3. Furthermore, it has been shown by Maxwell that,
among all uniformly magnetized finite bodies, the field within a body is uniform
as well, if and only if the body is an ellipsoid [16]. Consequently, in the sphere
of Fig. 2, which is a special case of an ellipsoid, the field is uniform and given
by H = H V = −M/3, whereas in all the other bodies of Fig. 2, the field is
nonuniform. Equations (3) and (4) stating that H is irrotational and B is solenoidal,
yield, for the nonuniform field outside the sphere of Fig. 2, the values H = 2M/3
and H = −M/3 at the positions ∗ and , respectively. For the magnetized spherical
shell of Fig. 2, the nonuniform field is H = −M and H = 0 at the positions •
and x, respectively; in the inner empty sphere, the field is H = 0. Interestingly,
for the cylinder in Fig. 2, also each direction represents a principal axis of the
demagnetization tensor because the cylinder axis is a principal axis, and, for the
length-to-diameter ratio of 0.906, the demagnetization factor N = 1/3 has been
obtained numerically [17].
The field created by a PM outside the magnet is known as the stray field, whereas
the field within it is known as the demagnetizing field [9] – because the volume
average of the latter is nonpositive. However, as it can be seen in the example of
Fig. 3, at certain positions in a uniformly magnetized body, the field created by that
body can also be positive (H is parallel to M). Even if the demagnetization factor N
of the spheroidal shell in Fig. 3 is unknown, the values of H at the special positions
in the figure can be easily obtained taking into account the two demagnetization
factors of the larger and the smaller ellipsoid, Na and Ni , respectively.
1378 K.-H. Müller et al.

Fig. 2 Examples of the


numerous bodies with a
demagnetization factor of
N = 1/3. The cylinder has a
length-to-diameter ratio of
0.906. The symbols ∗, , •,
and x are discussed in the text

Fig. 3 Uniformly
magnetized spheroidal shell
based on two ellipsoids with
the demagnetization factors
Na and Ni and values of H at
the indicated positions. N is
the (unknown)
demagnetization factor of the
shell

As it will be discussed in section “Hysteresis Curves and Magnetization Pro-


cesses of Permanent Magnets”, the macroscopic magnetic behavior of a PM is
largely governed by processes that take place in very small spatial regions of few
nanometers in size. The foregoing examples show that the field even in uniformly
magnetized bodies can locally deviate from the volume average H V . In bodies
with corners, edges, or spikes (Fig. 2), such deviations are rather strong. In this
case, the field H has also components perpendicular to M (even if M is parallel
to a principal axis of N ), and, in strictly formal terms, its magnitude diverges log-
arithmically with distance to the corners or spikes [18, 19]. However, for distances
not considerably larger than the lattice constants, those results of micromagnetic
calculations become meaningless, and, in real systems, the magnitude of the local
field will not exceed 2M [18]. The stray fields caused by nonuniformly magnetized
bodies are more difficult to capture and can well be larger than 2M. An interesting
example is the Halbach cylinder shown in Fig. 4 [20,21,9]. There, the magnetization
has a fixed magnitude M, but its direction (perpendicular to the cylinder axis)
changes continuously, and its angle is twice the azimuth angle of the position with
respect to the vertical axis. Inside a long Halbach cylinder, the field is uniform and
is given by

H = M ln(r2 /r1 ) (8)


28 Permanent Magnet Materials and Applications 1379

Fig. 4 (Left) Uniformly magnetized long hollow cylinder; (right) long Halbach cylinder; at the
position ∗, the field in the material has the same value as in the bore, H . Thick arrows indicate the
local direction of the magnetization M

where r1 and r2 are the radii of the inner and the outer cylinder, respectively. Hence,
for (r2 /r1 ) > 2.72, the field in the bore, H , will be larger than M. However,
such field concentration has its price as can be seen in Fig. 4: because the field is
irrotational, its value at the position ∗ in the material is also H , i.e., it is antiparallel
to M. Consequently, H can only increase with increasing r2 /r1 as long as the
coercivity of the material is larger than H given by (8). For comparison, Fig. 4 also
shows that inside a homogeneously magnetized hollow cylinder, the field is zero, as
in the hole of the shell of Fig. 2.

On the Thermodynamics of Magnets

One might expect that equilibrium thermodynamics as commonly used to describe


fluid or solid phases could simply be transferred to magnetic systems by adding the

applied magnetic field H to the intensive thermodynamic parameters of pressure p
and temperature T . However, due to the long-range character of the magnetostatic
interaction, this is not possible without restrictions, and, therefore, care is necessary.
Normally, in equilibrium, there is a thermodynamic limit, i.e., for a sufficiently
large sample, its macroscopic properties are independent of sample shape and
do not change if the sample is further enlarged. But the examples of uniformly
magnetized bodies discussed above show that the shape dependence of properties
does not disappear if the samples become very large. Furthermore, it has been shown
that easy-axis materials exhibit domain branching [22], a fractal domain structure
where the mean domain size increases without limit if the sample size increases. An
example of domain branching is shown in Fig. 5. Interestingly, as shown in [22], for
sufficiently large samples, the domain width at the sample surface is independent of
sample size and is given by the intrinsic length Ds


AK1
Ds ≈ 3 Dc with Dc = 72 lκ κ = 72 2
. (9)
μ0 Ms2
1380 K.-H. Müller et al.

Fig. 5 Branched domains in


Nd2 Fe14 B grains in a
Nd-Fe-B ingot at zero applied

field H . Pictures are taken
nearly parallel to the
tetragonal c-axis, close to the
sample surface, and in the
neighborhood of a
misoriented grain. (After
[22])

W.F. Brown has shown that, in thermal equilibrium, there is a certain critical
diameter Dc above which spherical samples consisting of easy-axis material are not
in a homogeneous state but are nonuniformly magnetized [23], and the value of
the critical single-domain particle diameter Dc in (9) had been estimated by Kittel
[24]. According to Coey (9), the value of Dc can be easily obtained from pictures
of domains on grains in, for example, as-cast material such as in Fig. 5.
Non-homogeneous equilibrium states of chemically homogeneous systems such

as the domain structure at zero external field, H = 0, in Fig. 5 are not uncommon
in thermodynamics and are examples of spontaneously broken symmetry, like the
coexistence of water and vapor of H2 O at fixed values of volume and temperature
[25]. In an approximation neglecting the interface energy and the stray field energy,
the two types of magnetic domains, that is, the brighter and darker areas in Fig. 5
with magnetizations antiparallel to each other and parallel to the easy axis (i.e.,
the crystallographic c-axis), can be considered as different thermodynamic phases

coexisting at H = 0 [26]. The detailed shape of the equilibrium domain structure

results from taking into account interface energy and stray field energy. At H = 0,
in thermodynamic equilibrium, the net magnetization of a PM is zero, and the field
created by it is very small. If the grain size of the easy-axis magnetic main phase in
the material is considerably larger than Dc , such demagnetized states are realized
by magnetic domains within the grains as can be seen in Fig. 5. The width of the
interface between those domains, the Bloch wall width, is given by

A
δw = π l κ = π (10)
K1

[27]. If the grain size in a PM material is close to Dc or even smaller, a classical


domain structure cannot be formed, and other types of equilibrium demagnetized
28 Permanent Magnet Materials and Applications 1381

Fig. 6 Interaction domains


visualized by a Kerr image
taken perpendicular to the
texture axis of a thermally
demagnetized melt-spun and
subsequently hot-deformed
Nd-Fe-B material, i.e.,
nominal c-axis is
perpendicular to the imaging
plane. The grain size is
≈400 nm

states will develop as, for example, interaction domains shown in Fig. 6. The width
of these domain-like patterns is considerably larger than the grain size, and they are
not limited by regions, such as Bloch walls, which are smooth and can be easily
mathematically described [28].

PM materials are developed to create stray fields H at zero applied fields H and

even for finite H antiparallel to H . Thus, as a matter of principle, magnetized
PM are in relatively stable non-equilibrium metastable states. These states and,
in particular, their volume average of magnetization MV depend not only on
 
the actual applied field H but also on the time dependence of H in the past.
This phenomenon is known as magnetic hysteresis. Usually only the component

of MV ≡ M parallel to H is considered. As can be seen in Fig. 7, in a typical

PM, the hysteretic behavior disappears for sufficiently large fields |H | where a

unique dependence M(H ) corresponding to thermal equilibrium is observed. For

values of (M, H ) within the major hysteresis loop which is obtained by starting

the measurements at large positive or negative H , an infinite number of different
 
M(H ) curves can be measured belonging to different magnetic histories {H } of
the magnet. In Fig. 7, besides the major hysteresis loop, the initial magnetization
curve is shown, which is just one of this infinite number of curves.
Due to their non-equilibrium character, the hysteretic states show small effects
of relaxation called magnetic viscosity or magnetic aftereffect or time lag in
magnetization as discovered by J.A. Ewing [29]. The result of a typical viscosity
experiment is shown in Fig. 8. For a large class of experiments, a logarithmic time
law

M(t) = M(0) + S ln(1 + t/t0 ) (11)

has been found where S is the magnetic viscosity and t0 is some reference time
depending on the investigated sample as well as on the measuring procedure. The
first theoretical analysis of magnetic viscosity was done by F. Preisach [30], who
1382 K.-H. Müller et al.

Fig. 7 Magnetization curves of a typical permanent magnet; thick blue, measured major hysteresis
loop with Mr as the remanence and Hc the coercivity; thin blue, initial magnetization curve (or
virgin curve) starting at the thermally demagnetized state; arrows indicate the direction of field
changes; dashed dark blue, states of thermal equilibrium

Fig. 8 Time dependence of


the magnetization of a
Nd4 Fe77 B19 permanent
magnet after the sample had
been saturated in a large
positive external magnetic
field and then experienced a
negative field jump. S and t0
are the fitting parameters of
(11). (Modified after [10,
Fig.1b])

suggested that it is based on thermal fluctuations and pointed out that the logarithmic
time dependence can be understood as a superposition of exponential relaxation
processes with a broad range of participating relaxation times. Later it was found
that the magnetic viscosity S does not go to zero but approaches a finite constant
value at low temperatures, and the reason for this observation is that quantum
tunneling of magnetization also results in magnetic viscosity, that is, in addition to
thermal fluctuations [31]. Preisach introduced a simple model connecting hysteresis
with viscosity. In this model, ferromagnetic materials are considered as collections
of bistable units each of them characterized by its individual rectangular non-
symmetric hysteresis loop (see Fig. 9). If the state of such an individual unit is on

its upper branch (js ) or its lower branch (−js ) and the applied field H is near the

critical value b − a or b + a, respectively, very small changes of H will result in
switching of that unit and, consequently, will contribute to an irreversible change of
the total magnetization M along macroscopic hysteresis curves of the material. In
such a critical state of the bistable unit, switching would also be caused by small
thermal fluctuations, thus contributing to the aftereffect, that is, to a spontaneous
28 Permanent Magnet Materials and Applications 1383

Fig. 9 Hysteresis loops of


two individual bistable units
of the Preisach model with j
as the magnetization of the
unit and the parameters js , a,
and b. (left) b > 0, a > b;
(right) b < 0, a < | b |.
(Modified after [10])


change of M at constant H . In recent years, Preisach-type models had a great
renaissance [32] and have been widely used to describe hysteresis as well as
relaxation phenomena.
In a thermodynamic description of uniformly magnetized materials, thermal

equilibrium is described by the Gibbs free energy G(p, T , H ), which is the
thermodynamic potential for given values of pressure, temperature, and external

field, p, T, and H , respectively, and equilibrium states as those represented by the
dashed dark-blue curve in Fig. 7 should be given by
 
 1 ∂G
M(H ) = −  (12)
μ0 V ∂H p,T

where V is the volume of the magnet. The Gibbs free energy G of a system in
the absence of an internal constraint is the minimum of the Gibbs free energy of a
manifold of constrained equilibrium states, G̃, [33]. This can be expressed by

 
G(p, T , H ) ≤ G̃(p, T , H ; λ) (13)

where λ is a parameter symbolically representing these constrained states. However,


the concept of constrained equilibrium is difficult to apply to magnetized states, and
in case of uniformly magnetized systems, Landau and Lifshitz had the ingenious
idea to use the magnetization itself as the internal parameter λ. Making plausible

assumptions on the function G̃(p, T , H ; M), they found what is known as the
Landau theory of phase transitions [34]. However, this is not sufficient to describe
the behavior of magnets because even in thermal equilibrium, the magnetization can
be nonuniform (as in Figs. 5 and 6), and, furthermore, metastable non-equilibrium
states have also to be taken into account. On the other hand, a general non-
equilibrium thermodynamics does not exist. The answer of micromagnetism to
these problems is to treat the metastable states like equilibrium states, that is, to
consider the relaxation times of the former to be infinite and to introduce a Landau-
type potential Ĝ for the manifold of arbitrary functions M(r) only constrained to a
constant magnitude Ms
1384 K.-H. Müller et al.

  
Ĝ(A, K, H ; [M(r )]) → min with |M(r )| = Ms (14)

where the square brackets indicate a non-local dependence of Ĝ on M(r) because


the stray field H generated by M(r) is governed by (3) and (4) and is non-locally

connected with M(r) as well. (Note that the prime at H in (14) has a different

meaning than that at r which represents all of the positions in the considered
body.) Consequently, there is no density of the potential Ĝ, depending only on local
quantities [35]. It is generally accepted that the global minima of (14) yield the
M(r) of thermal equilibrium, whereas the local minima represent the manifold
of metastable M(r) dependencies, of the considered magnet, where the spatial
averages MV are the values of magnetization measured on hysteresis curves (for
details, see Chapter "MIC"). It should be noted that in literature, the non-equilibrium
Landau-type expressions G̃ in (13) and Ĝ in (14) often are also termed Gibbs free
energy.

Hysteresis Curves and Magnetization Processes of Permanent


Magnets

The data characterizing the hysteretic behavior of a PM material are called extrinsic
properties because they are governed not only by the intrinsic magnetic properties
of the main phase in the material, which are in the simplest cases the quantities Ms ,
K1 , A, and Tc , but also by details of the microstructure of the material. One of the
important extrinsic properties is the remanent magnetization or simply remanence
Mr (see Figs. 7 and 10). Obviously an upper limit for the remanence of a material
is the spontaneous magnetization of its magnetic main phase, Mr ≤ Ms . The
difference between Mr and Ms is mainly caused by incomplete magnetic alignment
and by the presence of additional phases in the material.

Fig. 10 Permanent magnet with a square second-quadrant M − H major demagnetization curve


with Mr > Hc > Mr /2 and corresponding B − H curve. In that case, the (−BH )max value
is achieved at H = −Mr /2. (Without limiting the generality, the field H is thought to be the
demagnetizing field of ellipsoidal bodies)
28 Permanent Magnet Materials and Applications 1385

For an arbitrarily shaped body carrying a magnetization M(r), according to (3),


(4), and (5), it follows for the integral over the whole space

dV BH = 0 (15)
space

and consequently
  
dV BH = − dV BH = −μ0 dV H 2 = −2 Eout (16)
body out out

or

Eout
B(−H )body = 2 . (17)
V

Hence, the volume average of −BH within the magnetized body is twice the value
of the magnetostatic energy Eout created by that body outside of it, divided by its
volume. If M, B, and H are uniform in the body, averaging is not needed, and
the negative value of BH of the magnet is often slightly incorrectly termed its
energy product (or energy density). The second quadrant of a major hysteresis
loop like that in Fig. 7 can be realized by using the magnets’ own demagnetizing

field H instead of the external field H , for example, by using ellipsoids with one
of their principal axes parallel to the magnetization M. If now the corresponding
demagnetization factor N varies between 0 and 1 (long cigar to flat disc), the value
of B runs from Br = μ0 Mr to 0, representing the second quadrant of the major
B − H demagnetization curve (see Fig. 10). Obviously, on that path on the B − H
curve, (−BH ) starts and ends at the value 0 and achieves somewhere in between
its maximum value, which is the so-called maximum energy product (BH )max ,
one of the most important extrinsic magnetic properties of a PM material. In our
gedanken experiment with the varying ellipsoids, in case of the PM material of
Fig. 10, the demagnetization factor needed to achieve the (BH )max is N = 1/2, and
we have (−BH )max = μ0 Mr2 /4. If Hc in Fig. 10 were smaller than Mr /2, the value
of (−BH )max would be smaller than μ0 Mr2 /4 and would be limited by Hc as will
be discussed later (see (22)). It is obvious that these results can be generalized to the
case of an arbitrary major demagnetization curve:

(−BH )max ≤ μ0 Mr2 /4 ≤ μ0 Ms2 /4 (18)

The coercive field or coercivity Hc is a further, very important extrinsic property


of a PM material. For reasons of stability, in typical applications, Hc should be
considerably larger than limiting value, Mr /2, discussed above for the case of a
square demagnetization curve. The value of Mr can be considered as a good guide
value for the lower bound of coercivity needed in many cases of applications:
1386 K.-H. Müller et al.

Hc  Mr . (19)

To theoretically predict the value of Hc for a given material and to prepare


an appropriate microstructure in order to achieve large values of Hc are very
challenging problems in physics and material science of magnetism that have not
yet been really satisfactorily solved. W.F. Brown derived the famous formula

K1
Hc > 2 − NMs (20)
μ0 M s

as a lower bound for the coercivity (in terms of the external field as specified
in Fig. 7) of an ellipsoid with one of its principal axes, as characterized by the

demagnetization factor N, being parallel to the external field H ; and the ellipsoid
is assumed to consist of a homogeneous material characterized by a spontaneous
magnetization Ms , an exchange constant A, and the second-order anisotropy
constant K1 , and the presence of nonvanishing higher-order anisotropy constants
is not excluded [36]. It should be noted that (20) is independent of the sample
size! Although this inequality is a rigorous result of micromagnetism, the Hc values
measured on real specimens are orders of magnitudes smaller than the right-hand
side of (20) in most cases. Only in some rare cases of magnets obtained by great
efforts in industry and laboratories a coercivity could be achieved that exceeds
20 to 30 percent of the right-hand side of (20) [37]. These discrepancies, known
as “Brown’s paradox,” can only be explained by the fact that the assumptions
made by Brown are never met in real situations. Thus the samples are never fully
aligned ellipsoids, and, consequently, there are always components of the stray field

H perpendicular to H . Also, as discussed above, at corners or other irregularities
of the microstructure, the magnitude of H can be rather large. Furthermore, there
will always be local deviations from the regular chemical composition that result in
local magnetic softening. It is not yet well investigated how far these and further
irregularities may result in destabilization of the metastable magnetized states. At
this point, the well-known observation by J. C. Barbier [38] is helpful, who showed
that for a large class of magnetic materials, the coercivity Hc is strongly correlated
to the viscosity coefficient Sv (Fig. 11), which is connected to the viscosity in (11)
by S = χirr Sv , with χirr as the irreversible part of the susceptibility.
Furthermore Sv is related to the activation volume va in which the thermally
activated elementary processes of viscosity experiments are thought to take place
[39, 10]:

kB T
va = (21)
μ0 Sv Ms

where kB is the Boltzmann constant. Hence, va can be obtained from measurements


of S, χirr , and Ms and is found to be typically va ≈ δw 3 where δ
w is the
Bloch wall width (10), being often few nanometers in size in permanent magnets
[40, 41]. As a consequence of these results, it has been concluded that field-induced
28 Permanent Magnet Materials and Applications 1387

Fig. 11 Barbier plot


showing the relation between
coercivity Hc and viscosity
coefficient Sv for a large class
of magnetic materials. (After
[10])

elementary magnetization processes characterized by Hc are also initiated in small


volumes, typically va ≈ δw 3 in size [40]. Independent of this approach, coarse-

grained magnets with weakly interacting magnetic grains that are larger than Dc
from (9) have been successfully considered as nucleation-controlled magnets or
pinning-controlled magnets with large or small initial susceptibilities, respectively
(see Fig. 7) [42]. Prime examples for nucleation-controlled magnets are sintered
magnets based on SmCo5 and Nd2 Fe14 B, whereas magnets based on Sm2 Co17 are
usually considered as typical pinning-controlled magnets [42]. In both scenarios,
nucleation of reverse domains as well as pinning of domain walls, the elementary
processes are also assumed to take place in small spatial regions. Interestingly, it
has been found that the 2:17-type magnets could be also (and even better) described
as being controlled by interaction domains (similar as those in Fig. 6) where the
Sm2 Co17 cells of 50 to 100 nm in size have to be considered as weakly interacting
“grains” [43]. Generally the large class of fine-grained materials obtained, for
instance, by rapid quenching, mechanical milling, and HDDR processing, with
typical grain sizes smaller than Dc (often termed as nanocrystalline materials),
cannot be described as being nucleation controlled or pinning controlled but have to
be treated separately [44].

Permanent Magnet Materials

The first permanent magnet known to man was lodestone based on the common
mineral magnetite, Fe3 O4 . The ancient civilizations in Sumer, China, Europe, and
America were familiar with the natural magic of this magnet. It has not only
been a subject of curiosity, but already in ancient China, it has also been used
to manufacture a precursor of the modern compass [45]. At room temperature,
magnetite is cubic, and [111] are the easy axes of magnetization. Its spontaneous
1388 K.-H. Müller et al.

magnetization μ0 Ms is 0.6 T. In natural lodestones, the saturation magnetization


Msat is smaller than Ms due to the presence of other iron oxides, and the coercivity
μ0 Hc is typically 10 to 30 mT [46]. The coercivity mechanism is probably
pinning of domain walls at Fe2 O3 inclusions caused by oxidation of Fe3 O4 . Since
Hc  Ms , the (BH )max of lodestones is much smaller than the limit (18) and is
approximately given by

(BH )max ≈ μ0 Mr0 Hc (22)

where Mr0 is the ideal remanence of a non-textured (i.e., macroscopically isotropic)


material. The real remanence of a magnet, Mr , as defined in Fig. 7, can deviate from
Mr0 , in particular, due to effects of demagnetizing fields or effects of remanence
enhancement as will be discussed in section “Other Permanent Magnet Materials”.
In the case of [111] easy axes, Mr0 /Msat = 0.87 [47]. Assuming μ0 Hc = 10 mT
and μ0 Msat = 0.4 T, one gets (BH )max ≈ 3 kJ/m3 . However, for Hc  Ms , the
remanence of samples with a demagnetization factor N > Hc /Mr0 is approximately
given by

Mr ≈ Hc /N. (23)

It should be noted that the expression for Mr in (23) does not depend on Ms
and is usually below 0.1 T in lodestones and (BH )max < 1 kJ/m3 . In modern times,
lodestones have been replaced by artificial man-made magnets, and in particular,
in the twentieth century, (BH )max was improved enormously as can be seen in
Fig. 12.
In order to obtain a PM, the PM material has finally to be magnetized. For
the natural lodestones, the leading theory suggests that they were magnetized by
strong lightning strikes [48]. In the middle ages, in China as well as in Europe,
compass needles were made of steel and have been magnetized by touching them
with a lodestone or by quenching them from red heat in the Earth’s magnetic
field [9]. Since the nineteenth century, the steel magnets have been magnetized by
electromagnets, whereas since the twentieth century, magnetizing is mainly done
by using pulsed magnetic fields in the range of 1 to 10 Teslas with pulse lengths
of typically 0.5 to 1000 milliseconds, where capacitive discharge magnetizers are
mostly used. Because there are many technical difficulties in assembling devices like
motors, using magnetized magnets, post-assembly magnetizing has been developed,
where the magnets become magnetized in the final step of the device manufacturing
process. For this purpose, sophisticated magnetizing fixtures may have to be
designed [49]. A further special feature of modern magnetizers is that, if needed,
they can generate multipole magnetizations. Figure 13 shows a six-pole ring magnet
made by backward extrusion (at 750 ◦ C) of rapidly quenched Nd-Fe-B powder,
prepared to construct a brushless servomotor [50]. The hot deformation results in
a radial texture, where the local magnetically easy axis of the Nd2 Fe14 B phase in
the hollow cylinder is radially oriented, in contrast to the cases of hollow cylinders
28 Permanent Magnet Materials and Applications 1389

Fig. 12 Development of the “energy product” (BH )max in the twentieth century. Illustrating
the results, the different cylinders have a relative volume needed to store the same amount of
magnetostatic energy in the corresponding material. (Modified after [7])

Fig. 13 Field profile outside


a radially textured six-pole
Nd-Fe-B ring magnet. The
different colors represent
regions between isolines of
the field strength. Arrows
show local directions of the
field as well as the six radial
directions of magnetization of
the magnet pictured as the
black ring in the figure

in Fig. 4. The cylinder of Fig. 13 has a length of 25 mm, an outer diameter of about
33 mm, and an inner diameter of 27 mm. To obtain the six-pole field configuration
shown in the figure, a corresponding magnetizing coil for a pulse magnetizer needs
to be constructed.
In the following, we will discuss the properties of the most relevant materials
in the development shown in Fig. 12 – steel and alnico, hard ferrites, and rare-
earth material based on SmCo5 , Sm2 Co17 , and Nd2 Fe14 B. We will also consider
some further materials with remarkable properties and, in particular, those that are
considered as candidates for future PM materials.
1390 K.-H. Müller et al.

Hard Magnetic Steels and Alnico

The abovementioned compass needles made of steel are the first artificial permanent
magnets. Until the beginning of the twentieth century, the strongest magnets were

martensitic carbon steels, mainly consisting of the metastable martensite phase α -
Fe(C), stable phase ferrite α-Fe(C), residual austenite γ-Fe(C), and precipitations
of cementite Fe3 C. Their (BH )max was typically 3 kJ/m3 . Later, (BH )max was
increased by adding W, Cr, Mo, and/or Co. The strongest steel magnets have been
developed around 1920 in Japan. They contained the four elements Fe, C, Cr, and Co
and had a (BH )max of about 8 kJ/m3 [48]. Interestingly, the term “hard magnetic”
was originally used because the martensite steels are mechanically hard. Obviously,
the coercivity of steels not exceeding 10 mT is based on pinning of domain walls,
but the details of the pinning mechanism have not yet really been investigated. In
particular, the effect of the mentioned alloying elements and of the tetragonality of

a α -Fe(C) martensite (space group I4/mmm) are not yet understood.
In the 1930s, the research on hard magnetic steels led to the development of
a new iron-based series of compounds that were later named Alnico and had its
commercial success in the 1950s. This subject has been comprehensively reviewed
by several authors [51,52]. Besides the main components Fe, Co, Al, and Ni, Alnico
contains various further additives. The crucial difference between Alnico and hard
magnetic steels is that in Alnico, magnetic shape anisotropy is formed in addition
to the magnetocrystalline anisotropy where the latter is relatively weak in both types
of materials. Shape anisotropy stands for the fact that the magnetostatic energy of
an arbitrarily shaped uniformly magnetized body depends on the direction of its
magnetization M by

1
Em = μ0 V M N M (24)
2

where V is the body’s volume and N is its demagnetization tensor (6) [1]. In case
of axial symmetry, that is, if one of the principal axes of N has an eigenvalue
(demagnetization factor) of N and the other eigenvalues are 1/2 (1 − N ), one
can conclude from (24) that an effective anisotropy constant

(m) 1
K1 = μ0 (1 − 3N )M 2 (25)
4

can be introduced, which is formally equivalent to the previously discussed


anisotropy constant K1 in Equation (1). However, unlike K1 , it describes a
purely magnetostatic phenomenon. The microstructure of Alnico obtained by the
interesting thermodynamic process of spinodal decomposition [53] consists of thin
ferromagnetic needles rich in Fe and Co, dispersed in a nearly non-ferromagnetic
matrix rich in Al and Ni. In modern grades of Alnico (see, e.g., Table 1), these
needles are aligned, and N from (25) is the demagnetization factor of the entirety
28 Permanent Magnet Materials and Applications 1391

Table 1 Extrinsic magnetic properties of some magnet materials


Material Remanence Coercivity Energy product
μ0 Mr (T) μ0 Hc (T) (BH )max (kJ/m3 )
Co-steel 1.07 0.020 6 [54]
Alnico 5 1.28 0.064 44 [52]
1.35 0.074 50 [55]
Alnico 9 1.06 0.144 81 [52]
1.12 0.138 84 [55]

Fig. 14 (left) Circularly


magnetized torus with a
narrow gap; the symbol * is
explained in the text; (right)
typical shape of an Alnico
horseshoe magnet

of the needles. Obviously, the shape anisotropy is strong if the needles are thin and
long, and, consequently, N is small.
With the introduction of Alnico, it became possible, for the first time, to
replace electromagnets with permanent magnets in many devices such as generators,
motors, and microphones, which was a revolution in the use of permanent magnets.
The main advantage of Alnico is its structural as well as magnetic temperature
stability (its Curie temperature is 700 to 850 ◦ C [52]). Its biggest disadvantage is the
poor resistance to self-demagnetization due to its small coercivity Hc as compared
to remanent magnetization Mr (Table 1). The answer to this problem is the use
of horseshoe magnets that are known since the eighteenth century [9]. A typical
Alnico horseshoe magnet is shown in Fig. 14. The effect of its shape can be made
plausible as follows. Compact bodies have a demagnetization factor of typically
N ≈ 1/3, that is, if they are uniformly magnetized by a magnetization M, their
average demagnetizing field is about −M/3, which, according to Table 1, is much
too strong to maintain the magnetization M ≈ Mr of magnetic steel or Alnico.
Even in the case of a long rod with N  1, the demagnetizing field in the rod
close to its face side (position ∗ in Fig. 1) is −M/2, so even a long rod of Alnico
cannot be completely magnetized. If now the rod is imagined to be bent around
into a torus with a narrow gap (see Fig. 14, left), the demagnetizing field at the
position ∗ becomes very small, and the field within the gap will be H ≈ Mr . In
some approximation, a horseshoe magnet can be considered as a gapped ring as
shown in Fig. 14. Systems composed of magnets, air gaps of finite width, and soft-
magnetic segments are called magnetic circuits and are well described in many
publications [56,54,9]. Although Alnicos are still used in many special applications,
they have been gradually priced out of the market with the availability of hard
1392 K.-H. Müller et al.

ferrites and rare-earth magnets that can be made in any desired shape, without
having to resort to horseshoes to avoid self-demagnetization. Thus, Alnico magnets
saw no significant advancement in composition or processing since the 1970s,
until quite recently, stimulated by the rare-earth supply crisis and much advanced
characterization techniques [55, 57], new research started to better understand the
microstructure of Alnico and to improve its permanent magnet properties [55].

Hard Ferrite

The term ferrite is used for a wide class of iron-based oxides and, in metallurgy,
also for bcc α-Fe(C), whereas the term hard ferrite is usually confined to permanent
magnets based on the hexagonal compounds strontium ferrite, SrFe12 O19 (or
SrO·6Fe2 O3 ), and barium ferrite, BaFe12 O19 (or BaO·6Fe2 O3 ), discovered in 1951
in the Netherlands [58]. The dashed line in Fig. 12 represents Co-ferrite magnets
based on CoFe2 O4 that show (BH )max values of typically 1 kJ/m3 [59, 60].
CoFe2 O4 has the same spinel structure as Fe3 O4 . Intrinsic magnetic properties of
the hexagonal ferrites are shown in Table 2. As can be seen there, the anisotropy field
Ha of these compounds is considerably larger than their spontaneous magnetization
Ms . Thus, it was possible, for the first time, to prepare permanent magnets in any
desired shape without having to fear self-demagnetization. This is the main reason
why hard ferrites are much more common than Alnico although the energy density
(BH )max of the latter is considerably larger (see Fig. 12). The ferrite magnets based
on SrFe12 O19 are the most common, and according to [9], the following extrinsic
properties have been achieved for them: remanent magnetization μ0 Mr = 0.42 T,
coercivity (as defined in Fig. 7) μ0 Hc = 0.3 to 0.4 T, and maximum energy product
(BH )max = 34 kJ/m3 . Usually powders with particle sizes around 1 micron are
aligned in an external field and then sintered. A further advantage of the hard ferrites
in addition to their relatively large coercivity is their very low price as well as their
chemical stability and good corrosion resistance. The anisotropy field in Table 2 is
given by

K1
Ha = 2 . (26)
μ0 M s

Table 2 Curie temperature Tc and room-temperature values of other intrinsic magnetic properties
of some hexagonal ferrites. The data are taken from [9], with A, exchange constant; K1 , second-
order anisotropy constant; Ha , anisotropy field (26); Ms , spontaneous magnetization; κ, hardness
parameter (2); δw , Bloch wall width (10); and Dc , critical single-domain particle diameter (9)
Phase Tc A K1 μ0 Ha μ0 M s κ δw Dc
(K) (pJm−1 ) (MJm−3 ) (T) (T) (nm) (μm)
BaFe12 O19 742 6 0.33 1.7 0.48 1.3 13.6 0.56
SrFe12 O19 746 0.35 1.8 0.48
PbFe12 O19 725 0.25 1.5 0.41
28 Permanent Magnet Materials and Applications 1393

Fig. 15 Temperature
dependence of the anisotropy
fields of BaFe12 O19 and
SrFe12 O19 . (Reproduced
from [62], with the
permission of AIP
Publishing)

It should be noted that, in order to characterize the anisotropy of magnetic


phenomena, different characteristic fields can be and have been introduced taking
into account K1 as well as higher-order anisotropy constants. The meaning of Ha
  
from (26) is related to the slope of the magnetization curve, M⊥ (H ) = dM⊥ /dH ,
 
of a single crystal with its easy axis of magnetization perpendicular to H , at H = 0,

namely, Ha = Ms /M⊥ (0), and can be easily visualized by drawing a tangent line
  
on the M⊥ - vs. -H curve at H = 0. (Note that the prime at M has a different

meaning than that at H .) Interestingly, the anisotropy field of these 1:12:19 oxides
does not decrease monotonically if the temperature is increased but even slightly
increases at and above room temperature as can be seen in Fig. 15 [61, 62].
As a consequence of this behavior, the coercivity shows a stronger increase with
increasing temperature and an even more pronounced maximum than the anisotropy
field [63]. This can be understood considering the formula [64, 37]

Hc = α Ha − Neff Ms (27)

widely used to phenomenologically describe the coercivity by the two intrinsic


“fields” Ha and Ms (the latter representing the demagnetizing fields) where α and
Neff are dimensionless parameters and represent the effect of the microstructure
on coercivity. The effective demagnetization factor Neff is expected to be positive,
but, different from the demagnetization factors in (7), it is not necessarily < 1.
Equation (27) works well also for the hard ferrites whose coercivity is considered to
be governed by nucleation of reverse domains [65, 42]. The remarkable increase of
coercivity with increasing temperature can be explained by the fact that Ms in (27)
decreases monotonically, unlike Ha . Among all modifications, tried so far, of the
1:12:19 hard ferrites, the most promising is to partially substitute Co for Fe and, at
the same time, La for Sr (or Ba) because these new LaCo-type ferrite magnets show
significantly increased values of Ha and Hc , without a drop of Ms and Mr [66]. The
physics and manufacturing of hard ferrites are well described in many publications
[65, 67, 68, 69].
1394 K.-H. Müller et al.

Table 3 Curie temperature Tc and room-temperature values of further intrinsic magnetic proper-
ties of uniaxial ferromagnetic and ferrimagnetic phases commercially used for permanent magnets.
The data are taken from [9] with the exception of (1) , which is taken from [71], with A, exchange
constant; K1 , second-order anisotropy constant; Ha , anisotropy field (26); Ms , spontaneous
magnetization; κ, hardness parameter (2); δw , Bloch wall width (10); and Dc , critical single-
domain particle diameter (9)
Phase Crystal Tc A K1 μ0 Ha μ0 M s κ δw Dc
system (K) (pJm−1 ) (MJm−3 ) (T) (T) (nm) (μm)
BaFe12 O19 Hexagonal 740 6 0.33 1.7 0.48 1.3 13 0.6
SmCo5 Hexagonal 1020 12 17 40 1.07 4.3 2.6 1.2
Sm2 Co17 Trigonal 1190 16 4.2 6.5 1.3 1.8 5.7 0.5 (1)
Nd2 Fe14 B Tetragonal 585 8 4.9 6.7 1.61 2.5 4 0.2

Rare-Earth Magnets Based on SmCo5

An exciting idea is to combine the strong magnetic anisotropy of the rare-earth


elements which is caused by an efficient combination of crystalline electric fields
and spin-orbit interaction in these elements, with the large magnetization and high
Curie temperature of 3d elements, in particular, iron and cobalt. This idea was first
implemented by K. J. Strnat, G. Hoffer, and others in 1966, who found that SmCo5
has excellent intrinsic magnetic properties as can be seen in Table 3 [70]. It should
be noted that some of the quantities presented in this table show large scatter in data
reported in the literature. The main reasons for this are that the exchange constant A
is difficult to determine and, taking into account higher-order anisotropy constants,
the formulae for δw and Dc , (10) and (9), respectively, have to be modified [37].
As can be seen in Table 3, SmCo5 shows an extremely large magnetic anisotropy
characterized by an anisotropy field as large as 40 Tesla. Interestingly, this is caused
by both the Sm sublattice and the Co sublattice [9]. Due to the large value of K1 ≈
17 MJm−3 , the domain wall width δw is relatively small, but, nevertheless, it is at
least five times larger than the lattice constants in SmCo5 , and, consequently, SmCo5
is not a narrow domain-wall ferromagnet where the interaction of the domain walls
with the crystal lattice would contribute to coercivity as observed in certain rare-
earth compounds [72]. This is in accordance with the observation that in SmCo5
grains down to several microns in size, the domain walls are easily movable [73].
The world’s first rare-earth permanent magnet materials were those based on
SmCo5 . Typically, SmCo 1:5 magnets are produced using the powder metal-
lurgyroute, including a well-defined post-sintering heat treatment. Prior to sintering,
the magnetically easy crystallographic c-axes of the few micron-sized single-
domain SmCo5 particles are aligned magnetically in a green compact to induce
texture and thus large remanence Mr . These materials are typical examples of
nucleation-controlled permanent magnets, that is, their coercivity is limited by
nucleation of reverse magnetic domains at the grain boundaries [73]. However,
the nature of the grain boundaries of high-quality SmCo5 magnets and their
influence on the nucleation process are not yet fully understood and are still under
investigation [74]. The operating temperature of a PM material has not only to be
28 Permanent Magnet Materials and Applications 1395

significantly smaller than the Curie temperature; it also depends on how fast the
coercivity decreases with increasing temperature and up to which temperature the
microstructure of the material remains stable. Commercial SmCo5 magnets can be
used at temperatures up to 250 ◦ C or even higher because their intrinsic properties as
well as their microstructure are sufficiently stable. The values of remanence μ0 Mr ,
coercivity μ0 Hc , and energy product (BH )max achieved in them are 0.95 T, 2.5 T,
and 200 kJ/m3 , respectively. More detailed information on SmCo5 magnets can be
found in [75, 76] and [52].

Sm2 Co17 -Based Magnets

As can be seen in Table 3, Sm2 Co17 has a larger spontaneous magnetization Ms


than SmCo5 , and although Ha is smaller, it is still sufficiently large. This is the
main reason why larger (BH )max values have been achieved for PM materials
based on Sm2 Co17 (Fig. 12). Typical values of the extrinsic magnetic properties of
commercial 2:17 magnets are, at least, (BH )max = 250 kJ/m3 , μ0 Mr = 1.1 T, and
μ0 Hc = 1.6 T. These materials are also produced by powder metallurgy in order
to align the magnetically easy axes of the rhombohedral Sm2 Co17 phase. However,
the composition is nonstoichiometric, that is, as in the example of Fig. 16, the Sm
content is larger than 1/6, and further elements appear, in particular, Zr and Cu. The
microstructure of such magnets is shown in Fig. 16. It is obtained by a sophisticated
heat treatment procedure and consists of rhombohedral 2:17 cells (typically 100
nm in size) surrounded by few nm thick walls of 1:5 hexagonal structure that
are coherent with the 2:17 cells. It should be noted that, in spite of their finite
thickness, these walls are usually termed cell boundaries. Additionally, there are thin
lamellae (about 3 nm thick) perpendicular to the rhombohedral c-axis, consisting of
the Zr-rich Z-phase that has been described as Zr(Co,Fe)3 with the rhombohedral
Be3 Nb-type structure [77,78]. This structure has been confirmed by several authors,

Fig. 16 Transmission electron microscopy bright-field images of a sintered magnet of com-


position Sm(Co0.784 Fe0.100 Cu0.088 Zr0.028 )7.19 with (a) the nominal c-axis of the rhombohedral
Sm2 Fe17 main phase parallel and (b) perpendicular to the imaging plane (A indicates the 2:17
cells, B the 1:5-type cell boundaries, and C the Z-phase lamellae) [43]
1396 K.-H. Müller et al.

but its structure and chemistry are still somewhat controversial [43]. These lamellae
are considered to stabilize the cellular microstructure and, more importantly, to
modify the phase-ordering kinetics needed to obtain high-performance materials
[77]. The Z-phase has reported to be ferromagnetic with a Curie temperature of 473
K [79].
It has been proposed [80] that the coercivity of well-prepared 2:17-type magnets
is controlled by pinning of domain walls at the 1:5 cell boundaries. This is generally
accepted in literature and is in accordance with the magnetization behavior of these
magnets [81]. It should be noted that microstructures similar as that shown in
Fig. 16 and, consequently, high magnetic hardness of SmCo 2:17-type magnets are
achievable for various processing routes, namely, not only in sintered magnets but
also in ingot materials [78] and in melt-spun ribbons [82, 83]. In other words, the
actual grain size itself, pivotal for the magnetization reversal in many other PM
materials, plays only a minor role in 2:17 SmCo magnets. A crucial step in magnetic
hardening of this material is a final slow cooling from 850 ◦ C to 400 ◦ C (0.7 K/min),
which does not significantly change the cellular and lamellar microstructure but
results in a diffusion of copper from the 2:17 cell interior into the 1:5 cell boundary
phase (see Fig. 16) [84]. Furthermore, it has been shown that after the slow cooling,
the copper shows a strong maximum of its concentration in the middle of the cell
boundaries, probably much reducing the exchange interaction between neighboring
2:17 cells [82]. Consequently this material should behave like a well-textured fine-
grained material. This concept is supported by the domain patterns shown in Fig. 17
that are notably similar to those of Fig. 6 and can be considered as interaction
domains [43]. In Fig. 17, it can be seen that magnetic hardening of the 2:17 magnets
by chemical modification of the 1:5 cell boundaries not only changes domain

Fig. 17 Magnetic domain structure observed by magnetic force microscopy of a thermally demag-
netized Sm(Co0.784 Fe0.100 Cu0.088 Zr0.028 )7.19 magnet, with the nominal c-axis of the rhombohedral
Sm2 Co17 phase perpendicular to the imaging plane; (left), slowly cooled from 850 ◦ C to 400 ◦ C
(i.e., in the high coercivity state); (right), quenched from 850 ◦ C (low coercivity state). (Modified
after [43])
28 Permanent Magnet Materials and Applications 1397

wall pinning forces but changes the domain patterns themselves drastically. More
detailed information on Sm2 Co17 -type magnets and their history can be found in
[75, 76, 85, 86] and [52].

Nd2 Fe14 B-Based Magnets

Independent of the discovery of the ternary rare-earth transition metal compound


Nd2 Fe14 B in 1979 [87], the development of PM materials based on it started in
1983 using rapid solidification by melt spinning [88] and liquid-phase sintering
[89]. Since then, both techniques have been further developed by many groups, and,
in parallel, the physical properties of Nd2 Fe14 B have been successfully investigated
in detail. This extensive early work has been excellently reviewed in [90, 8, 91] and
[92]. Most of the commercial 2:14:1 magnets are produced by the sintering route,
for which (BH )max values as high as 451 kJ/m3 [93] and even 474 kJ/m3 [94] have
been achieved (Fig. 18). This is not far from the upper limit 516 kJ/m3 resulting
from (18). However, for most commercial sintered magnets, (BH )max is still below
400 kJ/m3 . The maximum achieved values of coercivity μ0 Hc and remanence μ0 Mr
are about 3.0 T [95] and 1.56 T [94], respectively, where, of course, both values
are not achieved for the same material. Weak points of Nd-Fe-B magnets are their
low corrosion resistance and their low Curie temperature Tc ≈ 315 ◦ C that limits the
operating temperature to about 150 ◦ C, which is not sufficient for many applications.
As can be seen in Fig. 18, this limit can be shifted up to about 250 ◦ C by partial
substitution of Nd by Dy (or, even more efficiently, Tb – not shown in the figure)
and of Fe by Co, however, at the expense of high performance at room temperature.
The corrosion behavior can be improved by small additions of Co, Nb, V, Mo,
Cu, and Ga that make the Nd-rich coatings of the single-crystalline 2:14:1 particles
in Nd-Fe-B magnets more noble [95]. Usually, these intergranular regions are
misleadingly termed grain boundaries (GBs). They consist of a solidified liquid
eutectic phase which makes liquid-phase sintering possible. The paramagnetic

Fig. 18 Dependence of
(BH )max on temperature for
sintered PM materials.
(Modified after [95])
1398 K.-H. Müller et al.

Nd-rich GBs are few nanometers thick and separate the 2:14:1 grains with respect to
exchange interaction, making the coercivity in sintered Nd-Fe-B magnets controlled
by nucleation of reverse magnetic domains [89, 96, 97]. Sagawa et al. found that the
coercivity of sintered Nd-Fe-B magnets can be considerably improved by a post-
sinter heat treatment at a temperature somewhere between 500 and 600 ◦ C, and they
explained this effect by a modification of the solidified liquid eutectic Nd-rich phase
[89]. If the magnets are doped by further elements like Cu and Al, the increase
of coercivity caused by such post-sinter annealing (PSA) can be remarkably large
[98]. Meanwhile, many investigations of this important effect have been published,
but it is not well understood. Certainly, solidification of thin layers of the liquid
eutectic phase will not result in the relatively coarse two-phase heterostructures
known for bulk eutectic systems. Nevertheless, the formation of non-homogeneous
concentration profiles, comparable to those observed in the cell boundaries of
heat-treated Sm2 Co17 -based magnets, as discussed in section “Sm2 Co17 -Based
Magnets”, cannot be excluded. Examples of mechanisms proposed in literature are
during PSA, the liquid eutectic phase dissolves surface irregularities at the 2:14:1
grains, consequently smoothing them [99]; PSA improves wetting of the 2:14:1
grains by the liquid eutectic phase [100]; PSA results in a beneficial redistribution
of Nd-rich material between the triple junctions and the GBs [101]; and even the
presence of a ferromagnetic phase in the GBs, caused by PSA, has been reported
[102, 103], what is in contrast to the commonly accepted paramagnetic GB model.
The extrinsic properties (BH )max , Hc , and Mr of sintered 2:14:1-based magnets
cannot be further improved much. However, in view of the expected increasing
demand in strong sintered Nd-Fe-B magnets, the process for producing them has
to be improved to get high-performance magnets at lower cost [104]. Two recent
developments are particularly interesting, the so-called pressless process (PLP) and
grain boundary diffusion (GBD) of Dy and Tb. In the case of PLP, the Nd-Fe-B alloy
powder is aligned in a pulsed field, but no heavy press machine and no dc electric
magnet are needed. Therefore, extremely fine and reactive powders with particle
sizes down to 1 μm, which requires milling in He, can be used and sintered without
serious oxidation, and, therefore, sufficiently large coercivity can be achieved for
magnets with reduced or even zero Dy content [105]. The Dy additions in Nd-
Fe-B magnets increase their anisotropy field Ha and, consequently, increase their
coercivity and their operating temperature. On the other hand, the Dy as well as
the Tb additions decrease the remanence Mr because the magnetic moments of
these heavy rare earths couple antiferromagnetically to the Fe moments, and, as
a further drawback, Dy and Tb are considerably more expensive than Nd. GBD is
successfully used to reduce the content of Dy needed in Nd-Fe-B magnets [106].
Thereby, small Dy-free magnets [107] or melt-spun ribbons [108] are coated with
a Dy-rich material, and in a subsequent heat treatment or during hot compaction, it
diffuses along grain boundaries into the interior of the magnet. There the Dy will
be mainly located in the intergranular regions and in regions of the 2:14:1 grains
that are close to the GBs, that is, at positions where nucleation of reverse magnetic
domains takes place. High-coercivity Nd-Fe-B without heavy rare-earth additions
could also be obtained by thick-film technology [109].
28 Permanent Magnet Materials and Applications 1399

Rapid solidification of Nd-Fe-B is the classical route to produce ribbons or pow-


ders with grain sizes in the submicron range, which are often called nanocrystalline
Nd-Fe-B powders. Nanocrystalline powders can also be produced by different
types of milling [114, 115] or the process of hydrogenation-disproportionation-
desorption-recombination (HDDR) [116, 117]. Normally, these techniques yield
powder particles of several microns in size or even larger, which consist of
isotropically distributed submicron-sized grains. As presented in Fig. 19, there are
a large variety of routes to produce isotropic (non-textured) as well as anisotropic
(more or less well-textured) magnets from fine-grained Nd-Fe-B powders [118,44].
For all materials shown in the bottom row of Fig. 19, remanences larger than
half of the spontaneous magnetization, Mr > 0.5 Ms , can be achieved by easy-
axis alignment (texture) or exchange coupling between submicron-sized grains
in isotropic materials (see section “Other Permanent Magnet Materials”). Well-
textured fine-grained Nd-Fe-B magnets can be obtained by deformation of the
rapidly solidified powders at a high temperature [119, 28, 120]. The underlying
mechanism of texture formation by hot deformation is a solution-precipitation
process [121]. The hot deformation is typically performed using die upsetting
or backward extrusion. The radially textured six-pole Nd-Fe-B ring magnet of
Fig. 13 is one example for hot-deformed melt-spun Nd-Fe-B. A special type of
hot deformation, namely, continuous rotary swaging, at elevated temperature has
been successfully applied directly to cast Nd-Fe-B alloy [122]. This resulted in
a remarkable increase of coercivity due to grain refinement during deformation.
However, grain alignment could not be achieved so far.

Fig. 19 Flowchart illustrating the principal processing routes of Nd-Fe-B permanent magnets
based on coarse-grained and nanocrystalline powders. (Modified after [44])
1400 K.-H. Müller et al.

Table 4 Comparison of room-temperature properties of selected fine-grained Nd-Fe-B materials


with those of hard ferrite and sintered Nd-Fe-B
Material μ0 M r (BH )max μ0 Hc
(T) (kJ/m3 ) (T)
Hard ferrite 0.42 34 0.35 [9]
Polymer-bonded melt-spun Nd-Fe-B (isotropic, MQ1-B) 0.69 80 0.9 [110]
Polymer-bonded HDDR Nd-Fe-B (isotropic) 0.52 37 0.6
Polymer-bonded HDDR Nd(Dy)-Fe(Co)-B (isotropic) 0.49 43 2 [111]
Hot-pressed melt-spun Nd-Fe-B (MQ2-E) 0.83 120 1.75 [110]
Commercial polymer-bonded Dy-free anisotropic 0.98 175 1.4 [112]
d-HDDR Nd-Fe-B [113]
Die-upset melt-spun Nd-Fe-B (MQ3-E) 1.31 335 1.6 [110]
Commercial sintered Nd-Fe-B 1.47 415 1.2 [95]
Advanced sintered Nd-Fe-B 1.519 451 0.98 [93]
Advanced sintered Nd-Fe-B 1.555 474 0.82 [94]

Another successfully established method is polymer bonding which has the


great advantage that complicated shapes and geometries can be prepared to near
net or net shape, using very low-cost production processes such as compression
molding and injection molding. Particularly important polymer-bonded magnets
are those made from anisotropic HDDR powder (Table 4). Originally it had been
reported that the multigrain particles of HDDR-processed Nd-Fe-B powders can be
anisotropic if the material is doped by certain additives such as Co, Ga, and Nb
[123, 124]. Later it could be shown that such doping elements are beneficial but
not necessary for inducing grain alignment within the powder particles obtained
by HDDR. Instead, under well-defined kinetic and thermodynamic conditions in
the “d-HDDR” process, tetragonal Fe2 B precipitates act as a memory, carrying the
information on the orientation of the original (not yet disproportionated) single-
crystalline powder particle to the nanocrystalline grains in the recombined material
[125, 126]. The great advantage of anisotropic HDDR Nd-Fe-B powder is that
it can be well aligned in a magnetic field and, unlike coarse-grained powder,
it has a sufficiently large coercivity and can be immediately used for polymer-
bonded magnets without any need of additional magnetic hardening. Finally, it was
demonstrated that the HDDR process is also a novel top-down processing route
to synthesize anisotropic nanocomposite Nd-Fe-B/α-Fe powders with substantial
coercivity [127].
A prospective preparation method of permanent magnets including most impor-
tantly Nd-Fe-B is additive manufacturing (AM), which generally benefits from
near-net-shape fabrication, i.e., the realization of complex shapes with minimum
material losses, the possibility of creating localized material states, or specifically
the graded properties at certain points of a magnetic body. AM can enable
also the combination of special magnetic, thermal, electrical, and mechanical
functionalities. Inspired from magnetocaloric cooling devices [128], one could,
for example, implement cooling channels inside a permanent magnet for high
28 Permanent Magnet Materials and Applications 1401

operation temperatures, which would be difficult to realize by conventional powder


metallurgical sintering [129]. Alternatively, one can calculate and print a magnet
body which provides a given magnetic field distribution [130, 131]. On the other
hand, magnetization for complex shapes is the main challenge, and texturing of
the grains has not been realized at all until today. Among the different types of
AM, the fused deposition modeling (FDM) also called fused filament fabrication
(FFF) or extrusion printing is a free space method, and large samples can be
realized by big area additive manufacturing [132]. In this method, wire-shaped
thermoplastic filaments or composite pellets are heated just above their softening
point and subsequently extruded by a moving head layer by layer onto a substrate
until the three-dimensional object is finished. Optimum flow properties are obtained
for spherical Nd-Fe-B particles as produced by inert gas atomization with typical
particle sizes of around 45 μm within a polyamide [130, 131]. However, up to
now, remanence is limited to 0.4 T in combination with a coercivity of 0.92 T due
to a reduced volumetric mass density compared to injection-molded specimens.
Note that for inert gas-atomized powder, alloying additions must be combined
with the Nd2 Fe14 B stoichiometric composition to avoid coarse under-quenched
microstructures. Alternative approaches for AM are the various powder bed methods
such as selective laser melting (SLM), selective laser sintering (SLS), binder jetting,
or electron beam melting (EBM). Among these methods, binder jetting is the
simplest one. Here, a print heat passes over the powder bed and selectively deposits
a polymer-binding agent to glue particles together layer by layer. After that, the
sample is heated for curing the binder. Unfavorable is the low volume density of
46% [133]. In terms of density, direct metal printing by SLM is most promising.
First results show the realization of a density of 92% with a volume fraction of the
Nd-Fe-B magnetic hard phase of around 70% and a grain size of 1 μm. Considering
the very low RE content of 19 wt.% given by the stoichiometry of the inert gas-
atomized powder, the obtained remanence of 0.6 T and coercivity of 0.87 T are
promising [129]. For further optimization, the complex peritectic phase formation
reaction in Nd-Fe-B is the major challenge, and the formation of Fe dendrites needs
to be avoided by cooling down the molten metal as quickly as possible. La-Fe-
Si, a magnetocaloric material, a similar phase decomposition reaction can occur
and is indeed triggered by SLM [128]; an additional annealing step after SLM
could however recover the required magnetic phases and thus functionality while
maintaining the complex geometry. Other general open issues are surface quality,
control of grain size, and functional and mechanic stability of the AM magnet.

Other Permanent Magnet Materials

In sections “Hard Magnetic Steels and Alnico”, “Hard Ferrite”, “Rare-Earth


Magnets Based on SmCo5 ”, “Sm2 Co17 -Based Magnets” and “Nd2 Fe14 B-Based
Magnets”, we discussed materials that are or have been widely used for permanent
magnets. However, there are many other materials that have been developed and
tested in laboratories and are produced only at small scale or have not been
1402 K.-H. Müller et al.

Table 5 Ferromagnetic and ferrimagnetic phases used to develop permanent magnet materials
[71, 117]
Phase Symmetry Tc μ0 M s μ0 Ha K1 δw Dc
(K) (T) (T) (MJ/m3 ) (nm) (nm)
α-Fe Cubic 1043 2.16
Fe65 Co35 Cubic 1170 2.43
Co Hexagonal 1390 1.81 0.76 0.43 14 68
Nd2 Fe14 B Tetragonal 585 1.61 6.7 4.9 3.9 ∼300
Pr2 Fe14 B Tetragonal 567 1.56 8.7 5 ∼4 ∼300
Dy2 Fe14 B Tetragonal 598 0.71 15.0 4.2 ∼4
Tb2 Fe14 B Tetragonal 620 0.70 22.0 6.4 ∼4
Nd2 Fe14 C Tetragonal 532 1.41 9.5 5.4 ∼4 ∼300
Y2 Fe14 B Tetragonal 575 1.32 1.05
Sm2 Fe17 N3 Rhombohedral 749 1.54 14 8.6 3.6 380
Sm2 Fe17 C3 Rhombohedral 668 1.45 16 7.4 ∼300
Sm2 Co17 Rhombohedral 1100 1.3 6.5 3.3 8.6 500
SmCo5 Hexagonal 993 1.05 40 17 3.7 1700
FePt Tetragonal 750 1.45 11.5 6.6 6.3 560
FePd Tetragonal 760 1.39 3.5 1.8 11.5 330
CoPt Tetragonal 720 1.0 12.3 4.9 7.4 1000
BaFe12 O19 Hexagonal 742 0.48 1.8 0.33 14 660
(Sm0.75 Zr0.25 )(Fe0.7 Co0.3 )10 Nx Hexagonal 877 1.7 7.7 ∼300
NdFe10 V2 Ny Tetragonal 743 1.11 10 ∼300
SmFe11 Ti Tetragonal 584 1.14 10.5 4.8 4 440
YFe11 Ti Tetragonal 525 1.07 0.85
Fe3 B Tetragonal 786 1.62 ∼0.4 -0.32 20
(Fe0.7 Co0.3 )2 B Tetragonal 937 1.31 1.0 0.42
YCo5 Hexagonal 987 1.07 6.5
La2 Co7 Hexagonal 479 0.7 6.7 1.4
MnAl Tetragonal 650 0.75 1.7
MnBi Hexagonal 633 0.73 0.9
Mn2 Ga Tetragonal >770 0.59 2.35
α -Fe16 N2 Tetragonal 810 2.41 1.0

developed to industrial maturity. The starting point of such developments is to


select ferromagnetic or ferrimagnetic phases that exhibit sufficiently high Curie
temperatures. Phases mostly used for this purpose are listed in Table 5. Iron and Fe-
Co alloys are not only the main phases in hard magnetic steels (see section “Hard
Magnetic Steels and Alnico”) but are also used in hard magnetic Fe-Cr-Co alloys
that are based on shape anisotropy similar to Alnico [52].
In 1988, Coehoorn et al. reported on an isotropic melt-spun Nd4 Fe78 B18 material
consisting of the nanocrystalline low-anisotropy phases α-Fe and Fe3 B together
with nanocrystalline Nd2 Fe14 B [135]. The most interesting properties of this
material are that, unlike the behavior of coarse-grained mixtures of these phases,
28 Permanent Magnet Materials and Applications 1403

its demagnetization curve does not contain any dip or step in the second quadrant
of the M-H’ plane and that the value of Mr is larger than expected (remanence
enhancement). Furthermore, the recoil curves are very steep, and the change in
magnetization along them is nearly reversible [134] as can be seen in Fig. 20.
These properties have been explained by exchange coupling of the different types of
small grains, and the materials showing such behavior have been called exchange
spring magnets [136]. This effect is closely related to the exchange coupling
between ferromagnetic and antiferromagnetic regions in a material, resulting in
exchange anisotropy [137]. Now many exchange spring-type materials are known
obtained by different preparation routes (see Fig. 19) and combining different types
of low-anisotropy and high-anisotropy phases. Usually, such materials are non-
textured, that is, the magnetically easy axes are randomly oriented [138, 139].
Among them, those based on melt-spun Fe/Nd2 Fe14 B are still most promising.
The typically achieved (BH )max is 180 kJ/m3 [139]. The weak point of spring
magnets is their low coercivity, μ0 Hc , usually not exceeding 0.5 T. A remarkable
increase of coercivity up to more than 1 Tesla could be achieved by doping melt-
spun Nd2 Fe14 B/α-Fe spring magnets with titanium [140]. An obvious idea is to
align the easy axes of the high-anisotropy grains in the fine-grained mixture. In a
much noted paper, Skomski and Coey showed, by micromagnetic calculations, that
in such a material, (BH )max values as high as 1 MJ/m3 should be achievable if the
grain size is a few nanometers [141]. To implement this practically has proved very
challenging and is still subject of ongoing studies [142]. One of the reasons for this
difficulty is the fact that it is very challenging to orient the nanostructured high-
anisotropy phase. Furthermore, the micromagnetic approach used in [141] may be
not sensitive enough to the details of the microstructure that govern the coercivity
mechanism in real nanostructured materials. A similar situation results in Brown’s
paradox (see section “Hysteresis Curves and Magnetization Processes of Permanent

Fig. 20 Demagnetization
curve and minor loops of the
spring magnet Nd4 Fe77 B19 .
(Modified after [134])
1404 K.-H. Müller et al.

Magnets”). At least, nanostructured multilayered films with well-oriented easy axes


and impressive magnetic properties have been obtained [143, 144, 142], which can
be considered as model systems for materials proposed in [141].
Remanence enhancement caused by exchange coupling also occurs in single-
phase nanocrystalline materials, as in melt-spun Nd-Fe-B, if the grains are suffi-
ciently small [145].
The tetragonal L10 phases FePt and CoPt have been successfully used to
fabricate permanent magnets with energy products up to 160 kJ/m3 [52, 7]. They
are ductile and chemically inert. However, they are extremely expensive and are
produced only in small quantities, mainly for medical implants.
The compounds Sm2 Fe17 N3 and Sm2 Fe17 C3 became promising candidates
for permanent magnets when it was discovered that more than two N or C atoms
per formula unit can be introduced in the rhombohedral compound Sm2 Fe17 , using
gas-solid reactions [146, 147]. The excellent intrinsic magnetic properties of both
interstitial compounds can be found in Table 5. The N or C atoms are on the
interstitial 9e site and form a triangle around the Sm atom on its 6c site, in a
plane perpendicular to the rhombohedral c-axis (see Fig. 21). The crystalline electric
field acting on the Sm atoms is dominated by contributions of the interstitial atoms,
resulting in a transition from easy-plane magnetic anisotropy (in Sm2 Fe17 ) to easy-
axis one. Such a transition had already been observed for Sm2 Fe17 Cx for samples
obtained by a melting preparation route [148]. However, for this technique, the C
concentration x is limited to 1.5, whereas x can be increased up to at least 2.3 by
using reactions with carbon-containing gas such as methane, acetylene, butane, etc.,
at around 550 ◦ C [149]. A complete filling of octahedral 9e sites would result in
x = 3. Nitrogen is introduced using NH3 or N2 . For Sm2 Fe17 Ny , an upper limit,
y ≤ 2.8, results from equilibrium thermodynamics [150]. It has been found that
only for R = Sm the interstitial modification of R2 Fe17 compounds results in the
easy-plane to easy-axis transition [151, 152]. To produce permanent magnets from
interstitially modified compounds, the most common routes are Zn bonding and
polymer bonding of Sm2 Fe17 Ny where (BH )max values around 100 kJ/m3 have
been reported [153,150,154]. To describe magnetization processes and coercivity in

Fig. 21 Interstitially
modified rhombohedral
Th2 Zn17 -type unit cell with
three formula units of
Sm2 Fe17 N3 or Sm2 Fe17 C3
28 Permanent Magnet Materials and Applications 1405

these materials, it has to be taken into account that the Sm2 Fe17 particles are never
fully nitrided which supports nucleation of reverse magnetic domains [155]. More
about the interstitially modified R2 Fe17 compounds and magnets made of them can
be found in [156, 153, 157, 158, 52, 150, 54, 9].

Resource Constraints

The search for new permanent magnet materials goes beyond modeling and
investigating the chemical and physical properties of a given candidate material.
A truly innovative, mass-market permanent magnet material (i) will have to be
produced at low cost on an industrial scale, using non-toxic raw materials and
processing routes; (ii) is expected to have an overall small ecological footprint;
and (iii) is based on low-cost raw materials with a stable supply worldwide. A
systematic study of these kinds of material properties would involve a resource
criticality assessment. It is a tool to identify and specify potentials and risks in
the use and processing of certain strategic raw materials [159, 160, 161]. A basic
approach to resource criticality is to specify two factors, the supply risk and
the economic importance of a strategic material [162, 163]. Beyond this two-
dimensional perspective, resource criticality can be described by factors such as
geological availability, geopolitical situation, economic development, substitution
potential, recycling potential, ecological footprint, and competing technologies in
the use of specific strategic materials [161].
Indeed, history shows that crucial technological breakthroughs in the devel-
opment of permanent magnets were driven by specific technological demands,
most importantly the compass, electric motors, aerospace engineering, magnetic
recording, and automation. Likewise, the success of certain material developments
was also influenced by resource constraints. As an example, the production and
use of ferrite magnets surpassed those of Alnicos in the 1960s, although ferrites
offer much lower energy densities (see Fig. 12). In fact, the low raw material and
production costs of ferrites were a key reason for their market breakthrough. Nd-Fe-
B permanent magnets were the answer to costly Sm-Co magnets developed in the
1960s and 1970s: in 1978/1979, the cobalt price exploded sixfold, due to increasing
global demands and the civil war in the Democratic Republic of the Congo, then
and now by far the main cobalt-producing country [164, 162]. In fact, for many
years, Nd-Fe-B was considered the perfect permanent magnet material. Apart
from its unsurpassed magnetic properties, the material is composed of elements
that are abundant in the Earth’s crust and that are comparatively inexpensive.
However, despite their abundancy, prices of rare-earth metals multiplied by an
order of magnitude within few months in 2010 and 2011 (Fig. 22). The diagram
shows that the world has become dependent on Chinese rare-earth production and
market policies: the price explosion correlates with a period of absolute Chinese
market dominance. A systematic criticality assessment provides a key perspective
to evaluate (i) why this price increase happened, (ii) if the global market will
experience a similar situation again within the next years to come, and (iii) if
1406 K.-H. Müller et al.

Fig. 22 Rare-earth crisis. World mine production of rare-earth oxides (according to [165]) versus
price development of Nd, Dy, Tb, and Pr (metal prices FOB China, according to [166]). Note:
There is a considerable uncertainty in estimating REO production outputs, at least in the order of
20 %

resource and material constraints might hinder the development of key future
technologies.
Geological availability: Rare earths occur in several places worldwide. They
are currently extracted, however, almost exclusively in China, apart from minor
production volumes in countries like Thailand, Russia, India, Australia, and the
USA (Fig. 22). US production was closed in 2015 and resumed in 2018. The
relative abundance of rare earths varies in relation to their atomic number: heavy
rare earths (Gd-Lu) are much less abundant than light rare earths (La-Sm). In
addition, rare earths with odd atomic numbers are less abundant than their adjacent
elements (Oddo-Harkins rule). Rare earths co-occur in natural deposits, due to their
chemical similarity. Thus, the mining and production of one rare-earth oxide will
always deliver a number of others. This phenomenon is called rare-earth balance
[167, 168, 169, 170]: today, the rare-earth market is mainly driven by the demand
for Nd, Pr, Dy, Tb, Y, and Eu for magnets and luminescent materials (Fig. 23).
However, producing 1 kg of Nd, the main deposits in China will also, on average,
produce about 2.5 kg of Ce and 1.5 kg of La, which are currently available in
excess [162, 161]. Taking a rare-earth balance perspective on the search for new
magnet materials clearly shows that a mass-market rare-earth permanent magnet
may only be based on La, Ce, Pr, and/or Nd as the main rare-earth metal component.
Furthermore, rare-earth minerals are mostly associated with other ore minerals, such
as those containing U and Th and/or Fe as it is the case of the Bayan Obo mine
(China). Economic and environmental issues in the extraction and handling of these
materials will ultimately also influence the economic success of the rare-earth metal
production.
28 Permanent Magnet Materials and Applications 1407

Fig. 23 Rare-earth balance – the oxides co-occur in nature. The diagram shows the average
distribution of the individual REE across 51 deposits worldwide, excluding Chinese deposits as
data is not accessible. (According to P. Christmann BRGM; see [162])

Geopolitical situation and economic development: The entire value chain


from mine to magnet is currently dominated by one, single economic and political
unit, which is China: more than 80 percent of rare-earth oxides as well as rare-
earth permanent magnets were produced in China in 2016 [162, 171]. This high
supplier country concentration has made industrial nations worldwide dependent
on Chinese policies. In 2011, rare-earth production was revived at Mountain Pass,
California, which once was the major rare-earth mine worldwide. Four years later,
the mine had to be closed again, due to the falling rare-earth prices, despite the
fact that – in the long term – a rising demand for rare-earth metals is forecast
and alternative, environmentally sustainable production routes are highly desired
by many stakeholders. A market failure like the 2010/2011 rare-earth crisis and
the risks in loosing technology leadership and direct control over REE primary
production were predicted by the US Geological Service in 2002, when the
Mountain Pass mine was closed for the first time [172,173]. Regarding supply chain
diversification, the global situation has not changed much since the rare earth crisis.
Substitution potential: Research into the substitution of Nd-Fe-B can be
considered on four different levels: (i) the substitution of individual elements
deemed critical; (ii) the substitution of an entire material; (iii) the development
of new, more material-efficient processing technologies; and (iv) the substitution
of entire technologies that use Nd-Fe-B. After 30 years of material optimization,
the substitution of individual elements in the Nd-Fe-B systems is unlikely to
bring additional benefit without the material losing performance. One approach
is a partial substitution of Nd by La and Ce, in order to produce less critical,
less expensive “rare-earth balance magnets” with better properties than ferrites
[174]. Alternative material systems that can compete with Nd-Fe-B, thus might
be considered as appropriate material substitutes, have not been discovered (sec-
tion “Future Permanent Magnet Materials”). This is different in terms of processing
technology: the grain boundary diffusion and the pressless process of Nd-Fe-B
(section “Nd2 Fe14 B-Based Magnets”) have both led to industrial breakthroughs in
the aftermath of the 2010/2011 global rare-earth crisis. Finally, a substitution of
Nd-Fe-B on the application level takes place in some areas, most importantly solid-
state drives for hard disk drives and induction motors for permanent magnet motors.
Nevertheless, this kind of substitution is very limited: in all kinds of motors that are
expected to provide high torque moments in small spaces (hybrid cars, bikes, fast-
moving robots) or – as it is the case of wind power – are expected to be technically
1408 K.-H. Müller et al.

robust and show a high energy efficiency, there are currently no alternatives to
motors containing Nd-Fe-B.
Critical competing technologies: It is meaningful to consider whether there are
other technologies that might compete with the magnet industry in the access and
long-term use of strategic metals. In the case of Nd and Pr, these are the glass, steel,
ceramic, and automobile (catalysts) industries. The absolute amounts used for these
fields of applications are, however, negligible. This is different for Tb: currently,
more than 70 percent of its global production is used in phosphors. However, this
is most likely to change, since LED lighting is now revolutionizing the market,
substituting traditional fluorescent lamps in many applications, which will result in
a lower demand for Tb in this application [175]. From a rare-earth balance point
of view, a growing availability of Tb would be welcomed by the magnet and motor
industry in order to make use of this heavy rare earth as an alloying agent in Nd-Fe-
B to increase magnetocrystalline anisotropy. Finally, Dy is almost exclusively used
in magnets.
Ecological factors: The mining and processing of rare-earth materials does
not necessarily come with a high environmental footprint. However, according to
literature, several mine operations, and particularly those in China, fail to apply
state-of-the-art environmental technology and management practices [173, 176].
This is, for example, related to insufficient control over filtering of acidic off-gases,
liquid discharges, mining waste management, and post-operational cleanup [173].
In addition, there are illegal mining practices with high impact. These malpractices
are of particular concern regarding the fact that rare earths often co-occur with Th
and U, and radioactivity should be carefully managed in mining, processing, and
waste management. Thus, today, rare-earth permanent magnets mostly come with
significant environmental footprints. This is challenging, particularly in the light of
their growing importance for e-mobility, wind power, air-conditioning, consumer
electronics, as well as robotics and industrial machinery. There is a rising awareness
for the environmental aspects of rare earth, and the Chinese government seems to
tighten rules and regulation for its rare-earth-producing companies [173].
Recycling potential: Industrial Nd-Fe-B recycling of end-of-life products is
still in its infancy. This is alarming considering the value of the material and
the environmental issues associated with its production. There are two major
approaches: first, to regain the metals or oxides using chemical extraction routes
(e.g., solvent extraction, selective precipitation) and, second, to pursue advanced
functional recycling routes that aim at re-processing the Nd-Fe-B alloys by casting
or powder metallurgical treatment, thus re-processing the material and thereby
preserving its functionality [181, 173, 182, 71]. The second kind of route is very
promising in terms of total energy and process costs. Nevertheless, particularly the
functional recycling approaches very much depend on well-defined feedstocks. This
is a major problem, since the extraction of scrap magnets from waste streams and
their characterization and classification are challenging.
Conclusions: The analysis shows that, today, Nd-Fe-B is a highly critical mate-
rial, although prices have not reached those levels of 2010/2011 again. We observe
a high supplier country concentration in raw materials production as well as in the
28 Permanent Magnet Materials and Applications 1409

Fig. 24 The demand for


Nd-Fe-B by different
applications in 2015.
Estimates according to
[171, 177, 178, 179, 180]. The
figures given by different
authors vary significantly.
Estimates for the total amount
of Nd-Fe-B produced
worldwide range from 80 to
140 kt in 2015. An average
value of 110 kt was taken for
this diagram

magnet industry itself; the material cannot be substituted in many key applications;
it has a significant ecological footprint; and the industrial-scale recycling of Nd-
Fe-B from end-of-life appliances is still in its infancy. The rare-earth and magnet
materials’ prices are not governed by a flexible, sustainable supply-demand balance
but by prices set by a Chinese planned economy and speculation. A way out of
the dilemma would be the development of equivalent substitute materials as well
as a diversification of the magnet supply chain (including recycling) by reviving
a magnet industry in several different places worldwide. Resource constraints
might once again turn out to be the catalyst for the development of new, stronger,
permanent magnet materials. At the same time, overcoming these constraints will
be crucial for the breakthrough of a number of emerging technologies, most
importantly e-mobility, large-scale wind power applications, robotics, and energy-
efficient industrial production machines (Fig. 24).

Future Permanent Magnet Materials

The energy product (BH )max = 474 kJ/m3 achieved for PM materials at room
temperature (Fig. 12) cannot be greatly improved for Nd-Fe-B materials because
1410 K.-H. Müller et al.

its value is close to its upper limit ≈ 510 kJ/m3 given by the relation (18). It is
unpredictable whether or not phases will be discovered that exhibit a spontaneous
magnetization exceeding that of Nd2 Fe14 B (1.61 T) and, at the same time, show
a sufficiently strong magnetic anisotropy and are sufficiently stable at processing
temperatures and operation temperatures at which magnets will be produced and
used, respectively. A possible candidate for this is the tetragonal compound α -
Fe16 N2 from Table 5. It was discovered by K.H. Jack who prepared it by a
quench-aging route starting with N-austenite γ-(Fe, N). The maximum N content
of γ and of N-martensite α that forms from it by quenching is considerably less
than 12.5 at%. Subsequent heat treatment promotes clustering of the interstitial
nitrogen in α and results in the two-phase system α -Fe8 N + α-Fe, and, at the same
time, α transforms into α where the N atoms occupy 2 of the 48 octahedral
interstices in a perfectly ordered manner: γ(Fe,N)→ α -(Fe, N)→ α -Fe8 N + α-
Fe → α -Fe16 N2 + α-Fe [183, 184]. Therefore, this conventional preparative route
will never produce more than 50 % α -Fe16 N2 . Further difficulties arise if the
processes are not fully completed, and, consequently, residual content of γ and
α further reduce the average magnetization of the material. However, in recent
years, α -Fe16 N2 nanoparticles were produced in gram amounts by a completely
different preparation route starting from iron oxides [185]. A new process for
the reduction of iron oxides at elevated pressures and high hydrogen pressure
and subsequent nitrogenation in an ammonia flow yielded fine, 99 % phase-pure
α -Fe16 N2 nanoparticles [186, 187, 188]. A serious problem is the fact that the
metastable phase α -Fe16 N2 decomposes at temperatures of at least ≈ 100 ◦ C [189],
and, therefore, an important goal is to stabilize it.
A more exotic compound is L10 -FeNi (tetrataenite) that is found in meteorites.
However, so far, thermodynamic and kinetic conditions for producing it are not well
known and may be difficult to implement [190]. The L10 -type lattice structure also
adopted by the equilibrium phases FePt, CoPt, and FePd is shown in Fig. 25. These
latter compounds can be obtained by annealing the quenched disordered cubic high-
temperature γ phases of A1-type structure, which is represented by the left part of
the figure if ac = c and both sublattices are statistically equally occupied. The L10 -
type structure (with a = c in the right-hand part of the figure) can also be obtained
by a tetragonal distortion of the cubic B1-type structure (with a = c). The following
intrinsic magnetic properties of L10 -FeNi, measured on meteoritic material, are

Fig. 25 Unit cell of the


L10 -FeNi lattice structure
(CuAu I prototype; space
group is P 4/mmm); (left)
representation with two
formula units; (right)
representation with one √
formula unit, with ac = a 2
28 Permanent Magnet Materials and Applications 1411

μ0 Ms = 1.47 T, μ0 Ha = 1.44 T, and Tc ≈ 550 ◦ C [191, 190]. It should be noted that


also L10 -FeCo has been investigated theoretically [192] and has been reported in
AuCu/FeCo nanocrystals [193]. These examples show that even the binary phases
that have already been exhaustively investigated can still cause surprise, but it may
be more promising to search for novel ternary or quaternary phases. Also hope is
not given up that the tetragonal ThMn12 -type rare-earth transition metal compounds
with their excellent intrinsic magnetic properties, such as NdFe10 V2 Ny , SmFe11 Ti,
and YFe11 Ti from Table 5, can be developed into practically relevant materials [92].
The binary 1:12 compounds do not exist, but the ThMn12 -type structure can be
formed with stabilizing elements.
As has been emphasized by Coey, in the near future, novel permanent magnets
will not only be developed with the aim to increase (BH )max but also to fill the gap
between the (BH )max values of hard ferrites (<40 KJ/m3 ) and of rare-earth magnets
(>200 KJ/m3 ), because such magnets with moderate performance will be needed
by numerous consumers [6, 194]. An interesting example is (Fex Co1−x )2 B that
shows uniaxial magnetic anisotropy with an anisotropy field μ0 Ha ≈ 1 T for x = 0.7,
although both parent compounds (x = 0 and x = 1) show easy-plane anisotropy
[195, 196]. Both parent compounds crystallize in the tetragonal C16-type structure
(space group I4/mcm, prototype CuAl2 ).
A particularly interesting group of materials are manganese compounds [194].
Manganese is abundant and cheap and forms alloys with high Curie temperatures. Its
atomic moments can be higher than that of Fe or Co. The difficulty is that Mn atoms
on high-moment sites tend to couple antiferromagnetically [194]. A promising
example of this group is the metastable ferromagnetic compound L10 -MnAl [197]
also known as τ phase (Table 5). Both elements, Mn and Al, are readily available
and cheap, and MnAl alloys are light and machinable. Unfortunately, stoichiometric
τ -MnAl is unstable. It has been stabilized by excess Mn as well as interstitial carbon
or partial substitution of Al by Ga [198]. Due to the presence of such additional
manganese as well as a certain degree of disorder in the occupancy of the two
sublattices (Fig. 25), which can never be totally avoided, some of the Mn atoms
reside on the Al sublattice. They couple antiferromagnetically to their Mn neighbors
in the Mn sublattice [199]. Furthermore, the magnitude of the magnetic moments
of the Mn atoms depends on their individual environments [200]. Consequently,
the spontaneous magnetization of τ -MnAl(C) samples depends on the content of
Mn and C as well as on the degree of disorder in the occupancy of the two
sublattices. As an example, μ0 Ms = 0.75 T has been reported for Mn54 Al44 C2 , and
μ0 Ha is 3.9 T and 5.3 T for Mn49 Al49 C2 and Mn54 Al44 C2 , respectively [201].
The typical quench-aging route for producing τ -MnAl(C) starts with the disordered
high-temperature ε-phase (space group P63/mmc, prototype Mg). The mechanism
for the τ -phase formation is discussed in literature, and there seems to be evidence
that both displacive and diffusional modes are included [202]. In any case, the
change in symmetry leads to a highly twinned material where three different types
of twins are usually present [203]. This makes it difficult to manufacture PM
materials with appropriate extrinsic magnetic properties, in particular, to achieve
well-aligned magnetically easy axes (i.e., the tetragonal c-axes). A degree of
1412 K.-H. Müller et al.

texture Mr /Ms as high as 0.85 has been achieved for hot-extruded MnAl(C)
[204]. Hot extrusion also results in an increase of coercivity due to grain refin-
ing. Extrinsic magnetic properties obtained for those magnets are μ0 Mr = 0.61 T,
μ0 Hc = 0.32 T, and (BH )max = 56 kJ/m3 [205, 206]. Later reported values for hot-
extruded gas-atomized MnAl(C) powders are μ0 Mr = 0.54 T, μ0 Hc = 0.25 T, and
(BH )max = 42 kJ/m3 [207]. A larger coercivity, μ0 Hc = 0.52 T, has been achieved
for annealed, mechanically milled Mn-Al-C powders [208, 209].
Among the Mn compounds, the tetragonal Heusler alloys Mn3 Ga and Mn2 Ga
as well as the L10 -phase MnGa are also under investigation [194, 210, 211, 212,
213]. However, this approach does not seem to be promising because the magnetic
properties of these materials are moderate and gallium is expensive.
The strategy for plugging the gap between high-performance rare-earth magnets
and hard ferrites should also include downgrading rare-earth magnets by substi-
tuting the expensive elements Dy, Tb, Nd, Sm by cheaper ones such as Ce and
La [214, 215]. Also, it should be emphasized that Alnico has a respectable energy
product (see Fig. 12) and a high Curie temperature. Drawbacks are its low coercivity
and the fact that Co is a strategic material. Recent investigations show that – now
that the microstructure of Alnico and its influence on Hc are better understood –
Hc can be possibly further improved [55]. An interesting question is whether PM
materials can be developed that are governed by shape anisotropy, comparable to
Alnico and FeCrCo, but contain only small or even zero amounts of cobalt. It cannot
be excluded that such low Hc magnets with moderate (BH )max will be used in the
future for energy-related applications.

Permanent Magnet Applications

In static applications, the working point P1 of a permanent magnet is fixed and,


in absence of other external fields defined as the intersection of the demagnetizing
branch of the magnet, obtained by a magnetic measurement in a closed circuit such
as in a permeagraph, with the load line or permeance line (Fig. 26a). The latter
is determined by the geometry of the magnet and accounts for the macroscopic
demagnetization field as described in section “Permanent Magnets and Physics
Behind Them”. For simple geometries such as a block [3] or a cylinder [216], the
demagnetizing factor N can be numerically calculated assuming a homogeneous
magnetization over the entire magnet giving the slope of the load line in the M-H
curve with −1/N and in the B-H curve −(1 − N)/N, respectively. If the sample
is measured in an open circuit, e.g., in a vibrating sample magnetometer (VSM),
the intrinsic hysteresis can be obtained by back-shearing the measured external
hysteresis with the slope of the load line in order to subtract out the stray fields in
the sample.
If the air gap between permanent magnets widens and narrows or the length of
a magnet shrinks or elongates with time, the working point P1 moves to P2 or vice
versa due to the change in the slope of the load line (Fig. 26b) [9]. In this dynamic
application with mechanical recoil, permanent magnets are displaced or rotated,
28 Permanent Magnet Materials and Applications 1413

Fig. 26 Demagnetizing curves showing the working points P1 and P2 for (a) a static application,
(b) a dynamic application with mechanical recoil, and (c) a dynamic application with active recoil,
modified after [9]. The dashed line P2 Q represents the stable trajectory with the slope μR known
as the recoil permeability. The blue square indicates the recoil product (BH )u

or soft iron is moved in the magnetic circuit. After several cycles, the working point
follows a stable trajectory, represented by the line P2 Q with the slope μR known
as the recoil permeability. The energy which can be used is described the recoil
product (BH )u as indicated by the blue rectangle in Fig. 26b and is much reduced
compared to (BH )max but is greatest for a broad square hysteresis loop and a recoil
permeability close to μ0 [217, 9].
In the presence of an external magnetic field generated by coils, the slope of the
load line in a fixed permanent magnet assembly remains but shifts parallel moving
the working point P1 to P2 (Fig. 26c). Such a case is called a dynamic application
with active recoil [217, 9]. In order to resist demagnetization, a high coercivity is
generally required for application. There are a number of additional properties of
permanent magnets that need to be considered: a high specific electrical resistivity
ρ in order to minimize eddy current losses, which are a major source of heating
the magnets; a square hysteresis loop to avoid hysteresis losses; as well as the
temperature coefficient of remanence and coercivity (α and β, respectively), which
are the key to evaluating magnetic performance at elevated operation temperatures
and ideally should be close to zero or even positive (as described in [7]). Typical
values are given in Table 6. Beware that the temperature coefficient β can only
describe a linear change in coercivity with increasing temperature, which is not
fulfilled for most of the materials. Furthermore, it must be considered that β also
depends on the degree of alignment and in the case of Nd-Fe-B magnets also on the
Dy content [222].

Permanent Magnet Assemblies and Their Dynamic Application with


Mechanical Recoil

Permanent magnets are widely used to generate static magnetic fields with relatively
high flux densities but without the need of a continuous energy input such as
electrical current as it is the case for solenoids. However, it is quite challenging to
1414 K.-H. Müller et al.

Table 6 Comparison of the specific electrical resistivity ρ and the temperature coefficients of
remanence α and coercivity β. The data are taken from (1) [9], (2) [218], (3) [89], (4) [219], (5) [220],
and (6) [221]
Material ρ α β
(μΩm) (%/K) (%/K)
SrFe12 O19 sintered(1) 108 −0.20 0.45
SrFe12 O19 polymer bonded(1) −0.02 0.45
Alnico 5 cast(1) 0.5 −0.02 0.03
SmCo5 sintered(1,2) 0.6 −0.04 −0.31
Sm2 Co17 sintered(1,2) 0.9 −0.03 −0.20
Nd2 Fe14 B sintered(1,3) 1.5 −0.13 −0.60
Nd2 Fe14 B die-upset(4,5) 1.2 −0.09 −0.60
Nd2 Fe14 B HDDR polymer bonded(1,6) 200 −0.10 −0.55

design permanent magnet assemblies with desired field uniformity or field gradient
within a specific macroscopic space. The major challenge is that a point dipole of
magnetic moment m measured in Am2 creates a nonuniform field which can be
expressed in polar coordinates [217]:

Hr = 2m cos θ/4π r 3 , Hθ = m sin θ/4π r 3 , Hφ = 0. (28)

Here, the direction and the magnitude of H depend on r and θ . By putting a


sufficient quantity of point dipoles in a row, the extended line dipole of magnetic
moment γ measured in Am creates a field:

Hr = γ cos θ/2π r 2 , Hθ = γ sin θ/2π r 2 , Hφ = 0. (29)

As Hr is now a factor of 2 smaller than for the point dipole, the magnitude of H

Hr2 + Hθ2 + HΦ2 is independent of θ , and the direction of H with respect to the
magnetization, η, is 2θ . By superimposing the magnetic fields of several line dipoles
or cylindrical segments with a homogeneous magnetization, a uniform field can be
created in a certain region. A prerequisite is that the field of one magnet segment
does not disturb the magnetization of the neighboring segments; in other words,
first the longitudinal and second the transverse susceptibility must ideally be zero.
For rare-earth or ferrite magnets, the first requirement is rather well fulfilled due
to a nearly rectangular M vs. H  hysteresis curve, and the second is only in the
order of Ms /Ha = 0.1. This rigidity of magnetization leads to the case that the
superposition of magnetic flux is linear and the magnetic material is magnetically
transparent, behaving like vacuum with permeability μ0 .
In a permanent magnet assembly of such cylindrical segments, the magnetic field
in a certain air gap, Hg , scales with the remanent magnetization Mr :

Hg = Ng Mr (30)
28 Permanent Magnet Materials and Applications 1415

with Ng as a scale-independent geometric factor of the magnetic circuit only


depending on relative dimensions instead of the absolute size of the circuit. If
Ng > 1, the magnetic field is concentrated. One important criterion of such
structures is the efficiency ε, which is defined as the ratio of the energy stored in
the air gap

Eg = 1/2 μ0 Hg2 d 3 r (31)

to that of the magnet



Em = 1/2 μ0 Mr2 d 3 r (32)

[9]. For a two-dimensional structure of infinite length, this leads to

ε = Ng2 Ag /Am (33)

with Ag and Am as the cross-sectional areas of the cavity and the magnet,
respectively. The theoretical upper limit of ε is 0.25. Therefore in reality a value of
0.1 is regarded already as efficient. Other criteria for a good structure are a minimum
of magnet material used to generate maximum uniform fields in a maximum sample
volume. More practically, the sample chamber should be easily accessible, and it
should be possible to vary the applied field between zero and maximum with small
amount of energy and torque.

Uniform Magnetic Fields


Uniform magnetic fields are generally desired to generate a well-defined torque Γ =
m × B on a magnetic moment m in order to align, for example, magnetic particles,
to measure hysteresis loops and Hall effects, or to create a magnetic resonance
image using the splitting of atomic or nuclear energy levels known as the Zeeman
splitting. Some sophisticated assemblies which generate a uniform magnetic field
are shown in Fig. 27. In Fig. 27a, wedge magnets with specific orientation can be
arranged according to Abele [223, 224] in √ a way that a maximum field of 0.293
Mr is realized for the dimensions r3 /r1 = 2 − 1 without producing stray fields
outside the assembly [9]. By nesting similar structures inside one another, multiples
of this field could be achieved. A different design, shown in Fig. 27b, uses a soft
iron yoke to close up a circuit around two permanent magnets generating fields
of typically 0.3 T in whole-body scanners. Both assemblies are widely used for
magnetic resonance imaging (MRI) to obtain a two-dimensional tomographic image
of solid objects [9]. For a full-body MRI system, approximately 1 t of Nd-Fe-B
magnets is needed [45]. A closed-up soft iron circuit can also be used to generate
a uniform magnetic field in a uniformly magnetized long hollow cylinder which is
1416 K.-H. Müller et al.

Fig. 27 Cross sections of some permanent magnet assemblies which create a uniform magnetic
field along the blue arrows (a) from wedge segments, (b) with soft iron return path, (c) from
Halbach array with soft iron pole shoes, (d) from segmented Halbach array, and (e) from a magnetic
mangle, respectively, modified after [217]. The black solid and the black dashed arrows indicate
the direction of magnetization in the permanent magnet as well as in the soft iron. The red dotted
arrows show the sense of rotation

magnetized as shown in Fig. 4 (left) and is fully covered with a thin iron sheet. In
the sense of a dynamic application, the sheet can be moved back and forth in order
to obtain zero or maximum field.
A famous structure is the Halbach array [21], which was already shown in
Fig. 4 but can also be modified by two pole shoes made of soft iron or FeCo
(permendur) in order to concentrate flux density to the saturation magnetization of
the pole shoe material of 2.15 and 2.45 T, respectively (Fig. 27c). This structure can
be considered as the three-dimensional extension of two one-sided flux tapes much
earlier proposed by Mallinson [20]. In this Halbach cylinder, the magnetization
rotates continuously and fulfills the same condition as for the line dipole, that is,
the direction of magnetization with respect to the vertical axis η at the position
P3 is two times the angular position θ , i.e., the angle between the vertical
axis and the position line (see Fig. 27c). For a Halbach cylinder with infinite
length, the geometric factor Ng is ln(r2 /r1 ) as shown in (8). For r2 /r1 = 2.2, the
efficiency is greatest with ε = 0.16 [9]. For a finite length L, (8) is modified
to (34) [225]:
28 Permanent Magnet Materials and Applications 1417


 
r2 z0
Hg (r = 0, z = 0) = Mr ⎝ln + 
r1 2 z02 + r12

⎛ ⎞⎞
z0 z0 +
z02 + r22
−  − ln ⎝  ⎠⎠ (34)
2 z0 + r 2
2 2 z0 + z0 + r 1
2 2

with z0 = L/2. If the Halbach cylinder is segmented into n trapezoidal magnets,


an extra term of sin(2π/n)/(2π/n) must be multiplied to the geometric factor
(Fig. 27d). This segmentation and the finite length mean that in a nested double
Halbach array, the field can never be cancelled out everywhere to zero [226].
Furthermore, a significant torque has to be applied in order to rotate the nested
Halbach arrays with respect to each other [226]. Whereas a Halbach array with 8
segments produces 90% of the flux density of a non-segmented Halbach cylinder, for
16 segments, one obtains 95% [226]. Large Halbach arrays weighting up to several
tonnes typically produce a field of about 1 T and were widely used for magnetic
annealing in the manufacturing of spin-valve sensors and to obtain a random access
memory on wafers up to 300 mm in diameter [9]. Furthermore, Halbach arrays are
becoming more and more important for magnetocaloric cooling devices, a new field
for dynamic application of permanent magnets [7, 227].
Halbach arrays are able to create flux densities much larger than the remanence of
the individual permanent magnets, but limitation is given (a) by the coercivity of the
permanent magnets, which must be higher than the desired flux density in the cavity,
and (b) by the logarithmic dependency in (8), which leads to an excessive increase
in the magnet material needed for gaining only minor increase in flux density. With
Halbach spheres, the field is a factor of 4/3 higher than for the Halbach cylinder, and
maximum fields of 4–5 T are realizable with Nd-Fe-B magnets, but there is limited
access to the sample chamber [228, 229]. The highest field of 5 T has been realized
in a tiny 0.15 mm air gap in a 120 mm Halbach sphere [229].
An alternative which benefits from the easy access is the magnetic mangle or
also referred to as magic mangle, which consists of cylinders whose magnetizations
are similarly arranged as in the segmented Halbach array (see Fig. 27a). If the rods
are touching each other, the field in the bore between n rods is [217]

Hg = (n/2)Mr sin2 (π/n). (35)

By rotating the rods alternately by ± 90◦ , a low-field configuration with zero field
at the center is obtained [226]. Arranging the magnetization in a ring results in
a larger low-field region, but this configuration cannot be achieved by continuous
rotation from the high-field state. The lower fields, and the much higher torque
needed for rotation, are drawbacks compared to the Halbach array, but lateral access
is convenient for a magnetic microscope stage [230].
1418 K.-H. Müller et al.

Nonuniform Magnetic Fields


Magnetic fields are commonly used to exert forces on charged particles due to the
Lorentz force F L = q (v × B). In uniform magnetic fields, this effect would cause
an electron to move in a helix in free space with the cyclotron frequency fc =
qB/2π me with me as the mass of the electron. In nonuniform magnetic fields, e.g.,
in an internal quadrupole field, as shown in Fig. 28a, an electron moving through
the cylinder would be focused in the middle where the field is zero, because any
deviation would cause an increased force back toward the center. This design was
the initial purpose of the Halbach array [21] and is nowadays used for cathode ray
tubes and charged particle beam control. Higher multipole fields are also possible.
In general, the orientation of the segments, η, fulfill the condition η = (1+(np /2)) θ
with np as the number of poles and θ as the angular position. Coaxial dipole rings
named magnetrons are often used in a sputter source near the targets to increase the
ionization rate in the plasma in order to yield a better sputter efficiency [217]. By
rotating the orientation of the segments which are marked in Fig. 28a with an ∗ by
90◦ creates an external quadrupole field without any field inside the bore following
the relation η = (1 − (np /2)) θ . This configuration is used as a rotor in electric
motors or generators, as explained in more detail in section “Dynamic Applications

Fig. 28 Cross sections of some permanent magnet assemblies which create a nonuniform
magnetic field along the blue arrows, modified after [217, 45]. (a) is an Halbach array with an
internal quadrupole field, (b) a magnetic mangle generating a field gradient, (c) a magnet for a
microwave travelling-wave tube, and (d) a wiggler magnet to generate electromagnetic radiation.
The * indicates the segments which have to be rotated to form an external quadrupole field
28 Permanent Magnet Materials and Applications 1419

with Active Recoil”. Unrolling such an assembly gives a one-side flux pattern which
is used for fridge magnets and other holding magnets made from bonded ferrites
[9]. Similar quadrupole fields are also possible with the magnetic mangle. Here also
very high field gradients are possible in order to exert forces on other magnets as
illustrated in Fig. 28b.
Another assembly of permanent magnets is shown in Fig. 28c. In this config-
uration, again a charged particle, e.g., an electron beam, is kept within a narrow
tube, but in this case, extra energy is coupled from a periodic axial cusp field which
generates longitudinal electron waves. This design is used in microwave power
tubes to amplify radiofrequency signals in the microwave range and to stabilize
molten metal flows. In synchrotron sources or free-electron lasers, a periodic
transverse field is generated from energetic electrons passing a wiggler structure
as shown in Fig. 28d in order to obtain an intense beam of hard radiation (UV and
X-ray).
One of the oldest applications for a permanent magnets is to attract other
permanent magnets or soft magnetic parts and to stick them together. The major
challenge is that permanent magnets are permanent and cannot be switched of in
order to release the attracted material. Switchable magnets which are widely used
for holding components in optical benches such as light microscopes, machine
tools, and magnetic clamps to fix signs and light objects to steel panels all use the
effect that moving a ferromagnetic part transverse to the magnetization direction of
a permanent magnet needs less force than longitudinal to it. In a closed circuit as
shown in Fig. 29, the energy in two infinite small air gaps with distance d is:

E = 2 × 1/2 μ0 Hg2 Ag d. (36)

Applying a force, F , the work done separating the segments is 2 F d which leads
to the maximum holding force Fh on the steel object in the on-position:

Fig. 29 Schematic illustration of (a) a switchable holding magnet and (b) electromagnetic
separation device, modified after [9]. The solid and the dashed arrows indicate the direction of
magnetization in the permanent magnet as well as in the soft iron. The dotted arrow shows the
sense of rotation
1420 K.-H. Müller et al.

μ0 Hg2 Ag
Fh = . (37)
2

Rotating the magnet to the off-position (mechanical recoil) closes the magnetic
circuit within the holding device, and the steel object can be lifted off. Other designs
of switching magnets use lateral movement of permanent magnet and soft iron
matrices against each other [9].
Nonuniform magnetic fields are furthermore used for mineral and metal
separation as shown in Fig. 29b. Hereby particles are moved over a fast rotating
wheel with an alternating permanent magnet configuration. If the particle has a
magnetic moment m, a volume V , and a susceptibility χ , a force Fm

F m = ∇(m · B) = (1/2) μ0 χ V ∇(H 2 ) (38)

is induced to it. For rare-earth permanent magnets, the magnetic field gradient
can reach values up to 100 T/m, and separation forces of order 108 Nm−3 are
possible which attract the particle to the spinning wheel [217]. Whereas the force
is proportional to the volume, the resulting trajectory depends on the force density
and is therefore independent of the particle size and only susceptibility-selective. If
the gradient is high enough, paramagnetic particles are also attracted, which is used
for separating paramagnetic deoxygenated red blood cells from whole blood or for
ferrofluid separation [9]. Using a fixed permanent magnet instead of a spinning
wheel is also thinkable; however, separation between metals and nonmetals is not
possible, because the eddy currents induced are not sufficient to significantly push
off the nonmagnetic metal particles from the magnet.
Other applications of permanent magnets are magnetic couplings, magnetic
gears, as well as magnetic bearings [217, 45]. The latter benefit from a cheap,
simple, and reliable design and can be arranged to bear a load linearly or radially.
Unfortunately, from Maxwell’s equation ∇ · B = 0, it is impossible to design a
static field configuration which allows a small magnet to be levitated at a fixed
point in space, i.e., to fulfill the condition that dB dB dB
dx , dy , and dz are all negative.
This impossibility known as Earnshaw’s theorem makes it necessarily to either use a
diamagnet or a superconducting permanent magnet [11,10] for levitation or to apply
mechanical or electromechanical forces in one dimension in order to prevent the
magnetic bearing from shifting or twisting [217]. Nevertheless, magnetic bearings
are widely used in flywheels for energy storage or in turbopumps, where a minimum
of friction is required. For magnetic levitation of a moving vehicle (MAGLEV), a
permanent magnet track provides either attraction to a soft iron sheet or repulsion
via eddy currents which is generally possible but very expensive.
A permanent magnet can also be used for magnetic sensor devices. Here the
change of magnetic flux between a stationary permanent magnet and a rotating
toothed wheel made of soft iron can be probed by a Hall sensor in order to detect
the sense of rotation of the wheel. Further applications are magnetically compen-
sated hinges and eddy current brakes in elevators as well as speedometers in
28 Permanent Magnet Materials and Applications 1421

automobiles. The latter systems must be as insensitive as possible to changes in


temperature; therefore, almost always Alnico magnets are used [45].

Dynamic Applications with Active Recoil

One major type of dynamic application with active recoil are actuators. Besides
the general meaning that actuators convert other types of energy into kinetic energy,
which a motor also does, actuators are generally understood as motors with limited
linear or angular displacement. In both cases, a coil acts with the field of a permanent
magnet. If the permanent magnet is stationary and the coil is moving back, it is
called a voice coil actuator. In this type of application, the coil moves in and out of
the magnetic flux Φ of the permanent magnet, which induces an electromagnetic
force (emf), that is, a voltage U in the coil due to Faraday’s law of induction
U = − dΦ dt .
In loudspeakers, which are one sort of voice coil actuators, the voice coil is
positioned in a small air gap of a magnetic circuit created by permanent ring magnets
mounted in soft iron rings (see Fig. 30a). Applying a current to the voice coil
moves the coil out (up) or in (down) again depending on whether the magnetic
field in the coil points down or upward. By applying an audio waveform, the
cone will reproduce the sound corresponding to pressure waves with an acoustic
power Pa depending on the field in the air gap Hg , which is maximized near
the (BH )max point. Therefore, loudspeakers made of Nd-Fe-B permanent magnets
can be designed much smaller than with ferrites, and the lightweight permanent
magnet can even be glued to the cone with a stationary drive coil. The reversed
effect is used in moving coil microphones, which are much more robust and cheap
compared to capacitor microphones. Furthermore, higher acoustic pressures are

Fig. 30 Schematic illustration of a voice coil actuator used in (a) loudspeakers and (b) a magnetic
hard disk drive, modified after [9]. The solid and the dashed arrows indicate the direction of
magnetization in the permanent magnet as well as in the soft iron. The dotted arrows show the
sense of rotation and linear movement, respectively
1422 K.-H. Müller et al.

detectable, and no voltage supply is necessary. The same effect is also used in voice
coil linear motors for any kind of linear motion where high precision and high
constant forces over the whole displacement length are necessary, as in moving coil
meters, pneumatic pumps, car door locks, and vibration control devices. Another
configuration is a flat voice coil used for accurate positioning of the read head
in a hard disk drive (see Fig. 30b). Here the coil moves transversely to the coil
axis,
rotating the read head around the pivot point. As access time scales with
1/ Hg , again Nd-Fe-B magnets with the highest energy densities are used. Similar
to microphones, a moving ferrite magnet in a stationary pickup coil converts the
oscillation signal of a magnetic steel string of an electric guitar into an electric
signal.
Another broad class of applications of permanent magnets are electric motors
or, using the reversed effect, generators. In motors, the coil of length L is
carrying a current I which, as a sum of all Lorentz forces of the electrons in
the winding, induces a net force F = L · I × B and rotates the rotor. In
comparison to induction-based asynchronous motors, permanent magnet-driven
synchronous motors are much more efficient, that is, they provide the same power
using smaller volume and mass [7]. A major drawback is that 70% of the material
costs is accounted for permanent magnets [231]. Thinking, for example, of a
3 MW direct-drive wind turbine, which typically uses 1.5 t of rare-earth permanent
magnets, or an electric motor and generator in a hybrid vehicle which needs 1.3 kg
of Nd-Dy-Fe-B magnets, this issue becomes important against the backdrop of
the resource criticality discussed in section “Resource Constraints”. Nevertheless,
electric motors are the major application of permanent magnet material in terms
of mass and quantity. The numerous variations of permanent magnet synchronous
machines can be distinguished by the position of the magnet, that is, whether the
permanent magnet is placed on the stator or on the rotator as shown in Fig. 31a, b.
In a two-pole brushed DC motor shown in Fig. 31a, the permanent magnet on
the stator creates a field which repulses or attracts the current-carrying windings of
the rotor. Mechanical commutation via brushes makes sure that the polarization of
the coil switches at each half turn and that the torque Γ on the rotor is always in the
same sense, leading to a rotation. The torque Γ is given by

Γ = C (U − Cω)/R (39)

with C = nw rw B L as the torque constant of the motor, rw the radius of the rotor
windings, nw the quantity of coils of length L, U the applied voltage, ω the angular
velocity, R the resistance of the windings, and Γ scales with the flux density B
produced by the magnet [9]. If the torque and the angular velocity are controlled by
modifying the applied voltage, it is a DC servomotor. In this case, monitoring of the
back emf or precise positioning is required. In order to eliminate the brushes which
are sources of sparking and wear, permanent magnets can also be placed on the rotor
and the windings on the stator (see Fig. 31b). In this case, electronic commutation
is possible either by applying an AC field or by a DC field. The latter requires
power controller in order to energize the windings in an appropriate sequence [9].
As synchronous motors lack starting torque until they reach synchronous rotation
28 Permanent Magnet Materials and Applications 1423

Fig. 31 Schematic illustration of (a) a two-pole DC electric motor with permanent magnets on the
stator and mechanical commutation (brushes) and (b) a brushless four-pole motor with a permanent
magnet on the rotor. The motor can be driven by a DC current combined with a power controller
(DC servo or spindle motor) or by AC current (synchronous motor). In the middle is a 13-bar
squirrel-cage winding using the principle of an asynchronous (induction) motor and its high startup
torque

speed, a squirrel-cage winding on the rotor is applied to maximize torque at startup.


A special type of DC servomotor is the spindle motor which is optimized for the
use in hard disk drives. Other variants of the DC motor are used for ventilators,
model airplanes, compressors, machine tools, and electric bicycles. Nowadays, in
terms of upscaling, synchronous permanent magnet motors with a power of 2 MW
are commercially available and typically applied in roller, press, and mill drives.
The inverse principle of a motor is used in a generator, for which the simplest
configuration is a bicycle dynamo. But much more important are wind generators as
one major source of renewable energy. Nowadays, wind turbines with a maximum
power output of several megawatts are standard, and even 10–20 MW has been
already realized [232, 7, 233, 234]. As the air-gap flux density of PM machines
is determined mainly by the quality of the permanent magnets utilized and does not
depend on the size of the machine, major limitation regarding upscaling is, apart
from the electromagnetically induced heating, given by the mechanical stress in the
rotor (for generators) or the rotor blades (for wind turbines) [233, 232]. Besides
the stress due to vibration or cogging [231], the tangential stress due to centrifugal
forces is dominant and proportional to the tangential rotor speed or circumference
speed, vt , which is defined as

vt = 2π rr fr (40)

with rr as the radius of the rotor and fr the rotational frequency of the rotor
[232]. The highest tangential rotor speed of 245 m/s at 120% overspeed has been
1424 K.-H. Müller et al.

realized for a large brown coal power plant induction-based generator at Lippendorf,
Germany, operating at 50 Hz with a rotor radius of only 1.3 m [233, 232]. For
PM motors, the embedded or surface-mounted magnets are more stress-sensitive.
Nevertheless, rotor speeds of 120–220 m/s are possible, for instance, using bonded
Nd-Fe-B ring magnets or Sm2 Co17 surface magnets fixed by a carbon fiber bandage
on a rotor with a radius of several cm [233]. For wind turbines, especially for off-
shore generators, the large rotor blades are the weakest points, and the rotational
speed must be low. In order to match the required power line frequency of 50 Hz
or 60 Hz in direct-drive wind turbines with maximum power output, the number
of poles has to be adjusted, or several generators have to be implemented in one
turbine [7]. Generally, the power output, P , of PM machines (motors or generators)
is indirectly proportional to fr [233, 232] following the relation

P ∼ 1/fr3.6 . (41)

From this point of view and considering the economical constraint (see above),
PM machines strongly dominate the field of machines below 500 kW [233], and
further upscaling of large off-shore wind turbines above the present maximum
of 20 MW will be challenging. For high-speed applications, e.g., miniature gas
turbines, mechanical factors such as stress and vibrations, rather than electromag-
netically induced heating, are the critical aspects which limit rotational frequency to
approximately 100.000 rpm at present [232].
A last important class of permanent magnet applications are stepping motors,
which can be considered as a discretized variant of the brushless DC motor. The
simplest configuration is the two-pole (bipolar) stepping (Lavet) motor called after
its inventor Marius Lavet (see Fig. 32a–d), which is nowadays used in most quartz
clocks. Here a permanent magnet is placed in a round cavity of a soft iron yoke
connected to a coil. Due to two small notches at the cavity, the magnetization
direction of the permanent magnet is energetically favored to lie transverse to the
notches (a). By applying a short current pulse along one direction, the permanent
magnet rotates to be consistent with the magnetization direction of the iron yoke (b).
After the pulse, the magnet rotates back to the nearest remanent position which is
now 180◦ off the initial remanent state (c). A second pulse in the opposite direction
rotates the permanent magnet further to be again consistent with the magnetization
direction of the iron yoke but now 180◦ off compared to the configuration after
the previous pulse (d). In this clever design, the permanent magnet, usually a small
Sm2 Co17 bonded magnet of a few milligrams in mass, rotates always in the same
direction. By increasing the number of notches and coils, a stepping motor which
rotates through a fixed angle of several degrees by driving a current through two
opposite windings is possible (see Fig. 32). In such a hybrid stepping motor shown
in this figure, the small step size of a reluctance stepper motor is combined with
the higher torque and the absence of a position loss in the zero current state of a
permanent magnet stepper motor. A sophisticated control is necessary to realize a
step-by-step rotation back and forth. Nowadays, hybrid stepping motors are almost
exclusively used for every kind of precise angular position control.
28 Permanent Magnet Materials and Applications 1425

Fig. 32 Schematic illustration of two-pole stepping (Lavet) motor in the (a) and (b) remanent
state and (c) and (d) applying a current pulse in the coil forward and backward. (e) shows a hybrid
stepping motor in the front and side view. The solid and the dashed arrows indicate the direction
of magnetization in the permanent magnets as well as in the soft iron. The dotted arrows show the
sense of rotation. (Modified after [9])

References
1. Brown, W.F.: Ferromagnetism. North-Holland Amsterdam (1962)
2. Brown, W.F.: Micromagnetics. Interscience Publishers (1963)
3. Aharoni, A.: Introduction to the Theory of Ferromagnetism. Oxford University Press (1996)
4. Heisenberg, W.: Zeitschrift für Physik 49, 619 (1928)
5. Bethe, H.: Annalen der Physik 3, 133 (1929)
6. Coey, J.M.D.: Scr. Mater. 67, 524 (2012)
7. Gutfleisch, O., Willard, M.A., Brück, E., Chen, C.H., Sankar, S.G., Liu, J.P.: Adv. Mater. 23,
821 (2011)
8. Herbst, J.F.: Rev. Mod. Phys. 63, 819 (1991)
9. Coey, J.M.D.: Magnetism and Magnetic Materials. Cambridge University Press (2010)
10. Müller, K.-H.: Magnetic viscosity. In: Buschow, K.H.J., Cahn, R.W., Flemings, M.C.,
Ilschner, B., Kramer, E., Mahajan, S. (eds.) The Encyclopedia of Materials: Science and
Technology, pp. 4997–5004. Elsevier Science Ltd (2001)
11. Müller, K.-H., Fuchs, G., Gutfleisch, G.: Permanent magnetism. In: Herlach, F., Miura, N.
(eds.) High Magnetic Fields: Science and Technology. World Scientific (2006)
12. Gruss, S., Fuchs, G., Krabbes, G., Verges, P., Stover, G., Müller, K.-H., Fink, J., Schultz, L.:
Appl. Phys. Lett. 79, 3131 (2001)
13. Durrell, J.H., Dennis, A.R., Jaroszynski, J., Ainslie, M.D., K.G.B. Palmer, Shi, Y.H.,
Campbell, A.M., Hull, J., Strasik, M., Hellstrom, E.E., Cardwell, D.A.: Supercond. Sci.
Technol. 27, 082001 (2014)
1426 K.-H. Müller et al.

14. COMSOL Inc. comsol multiphysics modeling software (2016)


15. Brown, W.F., Morrish, A.H.: Phys. Rev. 105, 1198 (1957)
16. Maxwell, J.C.: A Treatise on Electricity and Magnetism, vol. II. Clarendon. Oxford (1873)
17. Beleggia, M., De Graef, M., Millev, Y.T.: J. Phys. D Appl. Phys. 39, 891 (2006)
18. Zijlstra, H.: Permanent magnets – theory. In: Wohlfarth, E. (ed.) Handbook of Ferromagnetic
Materials 3. North Holland Publishing Company, Amsterdam pp. 37–105 (1982)
19. Thiaville, A., Tomáš, D., Miltat, J.: Physica status solidi (a) 170, 125 (1998)
20. Mallinson, J.C.: IEEE Trans. Magn. 9, 678 (1973)
21. Halbach, K.: Nucl. Inst. Methods 169, 1 (1980)
22. Hubert, A., Schäfer, R.: Magnetic Domains: The Analysis of Magnetic Microstructures.
Springer, Berlin/Heidelberg/New York (1998)
23. Brown, W.F.: Ann. N. Y. Acad. Sci. 147, 463 (1969)
24. Kittel, C.: Rev. Mod. Phys. 21, 541 (1949)
25. Anderson, P.W.: Basic Notions of Condensed Matter Physics. Benjamin/Cummings Pub. Co.,
Advanced Book Program (1984)
26. Landau, L.D., Lifshitz, E.: Course of Theoretical Physics: Vol.: 5: Statistical Physics.
Pergamon Press Oxford (1959)
27. Landau, L.D., Lifshitz, E.: Phys. Z. Sowjetunion 8, 101 (1935)
28. Khlopkov, K., Gutfleisch, O., Hinz, D., Müller, K.-H., Schultz, L.: J. Appl. Phys. 102, 023912
(2007)
29. Ewing, J.A.: Philos. Trans. R. Soc. Lond. 176, 523 (1885)
30. Preisach, F.: Zeitschrift für physik 94, 277 (1935)
31. Lambeck, M.: Barkhausen-Effekt und Nachwirkung in Ferromagnetika sowie analoge
Erscheinungen in der Festkörperphysik de Gruyter (1971)
32. Bertotti, G.: Hysteresis in Magnetism: for Physicists, Materials Scientists, and Engineers.
Academic Press (1998)
33. Callen, H.B.: Thermodynamics and an Introduction to Thermostatistics. Wiley (1985)
34. Landau, L.D., Lifshitz, E.: Course of Theoretical Physics: Vol. 8: Electrodynamics of
Continuous Medie. Pergamon Press (1960)
35. Brown, W.F.: J. Appl. Phys. 49, 1937 (1978)
36. Brown, W.F.: Rev. Mod. Phys. 17, 15 (1945)
37. Kronmüller, H., Fähnle, M.: Micromagnetism and the microstructure of ferromagnetic solids.
Cambridge University Press, Cambridge (2003)
38. Barbier, J.C.: Ann-Phys. 9, 84 (1954)
39. Wohlfarth, E.P.: J. Phys. F Met. Phys. 14, L155 (1984)
40. Givord, D., Tenaud, P., Viadieu, T.: IEEE Trans. Magn. 24, 1921 (1988)
41. Eckert, D., Müller, K.-H., Wendhausen, P.A.P., Wolf, M., Givord, D., Rossignol, M.F., Villas-
Boas, V.: IEEE Trans. Magn. 30, 574 (1994)
42. Livingston, J.D.: J. Appl. Phys. 52, 2544 (1981)
43. Gutfleisch, O., Müller, K.-H., Khlopkov, K., Wolf, M., Yan, A., Schäfer, R., Gemming, T.,
Schultz, L.: Acta Mater. 54, 997 (2006)
44. Gutfleisch, O., Bollero, A., Handstein, A., Hinz, D., Kirchner, A., Yan, A., Müller, K.-H.,
Schultz, L.: J. Magn. Magn. Mater. 242, 1277 (2002)
45. Du Tremolet, E., Gignoux, D., Schlenker, M. (eds.) Magnetism: I Fundamentals, II Materials
and Applications. Springer Science + Business Media (2005)
46. Wasilewski, P., Kletetschka, G.: Geophys. Res. Lett. 26, 2275 (1999)
47. Gans, R.: Annalen der Physik 407, 28 (1932)
48. Livingston, J.D.: Driving Force: The Natural Magic of Magnets. Harvard University Press
(1996)
49. Lv, Y., Wang, G., Li, L.: Rev. Sci. Instrum. 86, 034706 (2015)
50. Hinz, P., O’Beirne, D., Kirchner, A., Stass, W., Gutfleisch, O., V. Panchanathan, Müller, K.-
H., Schultz, L.: In: Kaneko, H., Homma, M., Okada, M. (eds.) REPM Proceedings of 16th
International Workshop on RE Magnets Applications. The Japan Institute of Metals, Sendai
(2000), pp. 1011–1020
28 Permanent Magnet Materials and Applications 1427

51. McCurrie, R.A.: The structure and properties of alnico permanent magnet alloys. In: E.P.
Wohlfarth (ed.) Handbook of Magnetic Materials, vol. 3. North-Holland Publishing Company,
p. 107 (1982)
52. Buschow, K.H.J.: Handb. Magn. Mater. 10, 463 (1997)
53. Cahn, J.W.: Acta Metall. 9, 795 (1961)
54. Skomski, R., Coey, J.M.D.: Permanent Magnetism. Institute of Physics Publication (1999)
55. Zhou, L., Miller, M.K., Lu, P., Ke, L., Skomski, R., Dillon, H., Xing, Q., Palasyuk, A.,
McCartney, M.R., Smith, D.J., et al.: Acta Mater. 74, 224 (2014)
56. Campbell, P.: Permanent Magnet Materials and their Application. Cambridge University Press
(1994)
57. Löwe, K., Dürrschnabel, M., Molina-Luna, L., Madugundo, R., Frincu, B., Kleebe, H.J.,
Gutfleisch, O.: Acta Mater. 407, 230 (2016)
58. Went, J.J., Rathenau, G.W., Gorter, E.W., Van Oosterhout, G.W.: Philips Tech. Rev. 13, 194
(1952)
59. Kato, Y., Takei, T.: J. Inst. Eletron. Eng. Jpn. 53, 408 (1933)
60. Heck, C.: Magnetic Materials and Their Applications. Butterworth (1974)
61. Jahn, L., Müller, H.G.: Phys. Status Solidi B 35, 723 (1969)
62. Shirk, B.T., Buessem, W.R.: J. Appl. Phys. 40, 1294 (1969)
63. Rathenau, G.W.: Rev. Mod. Phys. 25, 297 (1953)
64. Goodenough, J.B.: Phys. Rev. 95, 917 (1954)
65. Smit, J., Wijn, H.P.J.: Ferrites. Philips Technical Library: Eindhoven (1959)
66. Kools, F., Morel, A., Grössinger, R., Le Breton, J.M., Tenaud, P.: J. Magn. Magn. Mater. 242,
1270 (2002)
67. Kojima, H.: Fundamental properties of hexagonal ferrites with magnetoplumbite structure.
In: E. Wohlfarth (eds.) Handbook of Ferromagnetic Materials, vol. 3. Elsevier (1982), pp.
305–391
68. Stäblein, H.: Hard ferrites and plastoferrites. In: Handbook of Ferromagnetic Materials, vol. 3.
Elsevier (1982)
69. Kools, F., van der Valk, P.J.: Ferrites, in Kirk-Othmer Encyclopedia of Chemical Technology.
Wiley Online Library (2008)
70. Strnat, K., Hoffer, G., Olson, J., Ostertag, W., Becker, J.J.: J. Appl. Phys. 38, 1001 (1967)
71. Gauß, R., Diehl, O., Brouwer, E., Buckow, A., Güth, K., Gutfleisch, O.: Chemie Ingenieur
Technik 87, 1477 (2015)
72. Buschow, K.H.J., Brouha, M.: J. Appl. Phys. 47, 1653 (1976)
73. Livingston, J.D.: Phys. Status Solidi A 18, 579 (1973)
74. De Campos, M.F., Okumura, H., Hadjipanayis, G.C., Rodrigues, D., Landgraf, F., J.G., Neiva,
A.C., Romero, S.A., Missell, F.P.: J. Alloys Compd. 368, 304 (2004)
75. Strnat, J.K.: Rare earth–cobalt permanent magnets. In: Wohlfarth, E.P., K.H.J. Buschow (eds.)
Ferromagnetic Materials–a Handbook on the Properties of Magnetically Ordered Substances,
vol. 4. North-Holland Physics Publishing, pp. 131–209 (1988)
76. Strnat, K.J., Strnat, R.M.W.: J. Magn. Magn. Mater. 100, 38 (1991)
77. Rabenberg, L., Mishra, G., Thomas, R.K.: In: (ed.) Fidler, J. Proceedings 3rd International
Symposium Magnetic Anisotropy and Coerivity in Rare Earth-Transition Metal Alloys,
Austria (1982), pp. 599–608
78. Tang, W., Zhang, Y., Hadjipanayis, G.C.: J. Appl. Phys. 87, 399 (2000)
79. Katter, J.M., Weber, J., Assmus, W., Schrey, P., Rodewald, W.: IEEE Trans. Magn. 32, 408
(1996)
80. Livingston, J.D., Martin, D.L.: J. Appl. Phys. 48, 1350 (1977)
81. Durst, K.-D., Kronmüller, H., Ervens, W.: Phys. status solidi (a) 108, 705 (1988)
82. Yan, A., Gutfleisch, O., Gemming, T., Müller, K.-H.: Appl. Phys. Lett. 83, 2208 (2003)
83. Yan, A., Gutfleisch, O., Handstein, A., Gemming, T., Müller, K.-H.: J. Appl. Phys. 93, 7975
(2003)
84. Hadjipanayis, G.C., Tang, W., Zhang, Y., Chui, S.T., Liu, J.F., Chen, C., Kronmüller, H.: IEEE
Trans. Magn. 36, 3382 (2000)
1428 K.-H. Müller et al.

85. Hadjipanayis, G.C.: In: J.M.D. Coey (ed.). Ch. 6 in: Rare-earth iron permanent magnets, pp.
286–335. Oxford University Press (1996)
86. Dürrschnabel, M., Yi, M., Uestuener, K., Liesegang, M., Katter, M., Kleebe, H.-J., Xu, B.,
Gutfleisch, O., Molina-Luna, L.: Nat. Commun. 8:54 (2017)
87. Chaban, N.F., Kuz’ma, Y.B., Bilonizhko, N.S., Kachmar, O.O., Petriv, N.V.: Dopovidi Akad.
Nauk Ukr. RSR, Ser. A 873–875 (1979)
88. Croat, J.J., Herbst, J.F., Lee, R.W., Pinkerton, F.E.: J. Appl. Phys. 55, 2078 (1984)
89. Sagawa, M., Fujimura, S., Togawa, N., Yamamoto, H., Matsuura, Y.: J. Appl. Phys. 55, 2083
(1984)
90. Sagawa, M., Hirosawa, S., Yamamoto, H., Fujimura, S., Matsuura, Y.: Jpn. J. Appl. Phys. 26,
2785 (1987)
91. Herbst, J.F., Croat, J.J.: J. Magn. Magn. Mater. 100, 57 (1991)
92. Sepehri-Amin, H., Hirosawa, S., Hono, K.: Ch. 4 in Handbook of Magnetic Materials Vol.
27, E. Brück. Elsevier (2018)
93. Rodewald, W., Wall, B., Katter, M., Uestuener, K.: IEEE Trans. Magn. 38, 2955 (2002)
94. Harimoto, D., Matsura, Y.: Hitachi Met. Tech. Rev 23, 69 (2007)
95. Rodewald, W., Katter, M.: In: Dempsey, N.M., de Rango, P. (eds.) 18th International
Workshop on High Performance Magnets and Their Applications, pp. 52–63 Annecy, (2004)
96. Livingston, J.D.: In: K.J. Strnat (ed.) 8th International Workshop on Rare Earth Magnets and
Their Applications University of Dayton (1985), p. 423
97. Durst, K.-D., Kronmüller, H.: In: Proceedings 8th International Workshop on Rare-Earth
Magnets and Their Applications and the Fourth International Symposium on Magnetic
Anisotropy and Coercivity in Rare Earth–Transition Metal Alloys, pp. 725–735 (1985)
98. Eckert, D., Müller, K.-H., Wolf, M., Rodewald, W., Wall, B.: IEEE Trans. Magn. 29, 2755
(1993)
99. Ni, J., Ma, T., Yan, M.: J. Magn. Magn. Mater. 323, 2549 (2011)
100. Straumal, B.B., Mazilkin, A.A., Protasova, S.G., Gusak, A.M., Bulatov, M.F., Straumal, A.B.,
Baretzky, B.: Rev. Adv. Mater. Sci 38, 17 (2014)
101. Woodcock, T.G., Bittner, F., Mix, T., Müller, K.-H., Sawatzki, S., Gutfleisch, O.: J. Magn.
Magn. Mater. 360, 157 (2014)
102. Murakami, Y., Tanigaki, T., Sasaki, T.T., Takeno, Y., Park, H.S., T. Matsuda, Ohkubo, T.,
Hono, K., Shindo, D.: Acta Mater. 71, 370 (2014)
103. Sepehri-Amin, H., Ohkubo, T., Shima, T., Hono, K.: Acta Mater. 60, 819 (2012)
104. Sagawa, M.: In: S. Kobe, P.J. McGuiness (ed.) Proceedings of 21st International Workshop on
Rare Earth Permanent Magnets and Their Applications. Bled, Slovenia (2010), pp. 183–410
105. Sagawa, M., Une, Y.: In: D. Niarchos (ed.) Proceedings of 20th International Workshop on
Rare Earth Permanent Magnets and Their Applications. Knossos, Crete (2008), p. 103
106. Park, K.T., Hiraga, K., Sagawa, M.: In: Kaneko, H., Homma, M., Okada, M. (eds.) REPM
Proceedings of 16th International Workshop on RE Magnets and their Applications. The
Japan Institute of Metals, Sendai, pp. 257–264 (2000)
107. Löewe, K., Brombacher, C., Katter, M., Gutfleisch, O.: Acta Mater. 83, 248 (2015)
108. Sawatzki, S., Kübel, C., Ener, S., Gutfleisch, O.: Acta Mater. 115, 354 (2016)
109. Dempsey, N.M., Woodcock, T.G., Sepehri-Amin, H., Zhang, Y., Kennedy, H., Givord, D.,
Hono, K., Gutfleisch, O.: Acta Mater. 61, 4920 (2013)
110. Brown, D., Ma, B.M., Chen, Z.: J. Magn. Magn. Mater. 248, 432 (2002)
111. Chen, W., Gao, R.W., Zhu, M.G., Pan, W., Li, W., Li, X.M., Han, G.B., Feng, W.C., Wang,
B.: J. Magn. Magn. Mater. 261, 222 (2003)
112. Honkura, Y., Mishima, C., Hamada, N., Mitarai, H.: In: Proceedings of the 17th Internat.
Workshop on Rare-Earth Magnets and Their Applications (2002), pp. 52–61
113. Honkura, Y.: In: International Symposium Magn. Magnetic Materials, Current Resoultion
Trends Magnetic Materials, pp. 95–95 (2013)
114. Schultz, L., Wecker, J., Hellstern, E.: J. Appl. Phys. 61, 3583 (1987)
115. Crespo, P., Neu, V., Schultz, L.: J. Phys. D. Appl. Phys. 30, 2298 (1997)
116. Takeshita, T., Nakayama, R.: In: Proceedings of 10th International Workshop on Rare Earth
Permanent Magnets and Their Applications Kyoto, pp. 551–557 (1989)
28 Permanent Magnet Materials and Applications 1429

117. McGuiness, P.J., Zhang, X.J., Yin, X.J., Harris, I.R.: J. Less-Common Met. 158, 359 (1990)
118. Gutfleisch, O.: J. Phys. D. Appl. Phys. 33, R157 (2000)
119. Lee, R.: Appl. Phys. Lett. 46, 790 (1985)
120. Dirba, I., Sawatzki, S., Gutfleisch, O.: J. Alloys Comp. 589, 301 (2014)
121. Grünberger, W., Hinz, D., Kirchner, A., Müller, K.-H., Schultz, L.: J. Alloys Compd. 257, 293
(1997)
122. Chi, L., Wießner, F., Gröb, T., Müller, C., Durst, K., Bruder, E., Sawatzki, S., Löwe, K.,
Gassmann, J., Gutfleisch, O., Groche, P.: J. Magn. Magn. Mater. 490, 165405 (2019)
123. Takeshita, T., Nakayama, R.: In: S. Sankar (ed.) Proceedings of 11th International Workshop
on Rare Earth Permanent Magnets and Their Applications. Pittsburgh, pp. 49–52 (1990)
124. Takeshita, T., Nakayama, R.: In: Proceedings of 12th International Workshop on Rare Earth
Permanent Magnets and Their Applications, p. 670 Canberra (1992)
125. Gutfleisch, O., Khlopkov, K., Teresiak, A., Müller, K.-H., Drazic, G., Mishima, C., Honkura,
Y.: IEEE Trans. Magn. 39, 2926 (2003)
126. Honkura, Y., Mishima, C., Hamada, N., Drazic, G., Gutfleisch, O.: J. Magn. Magn. Mater.
290, 1282 (2005)
127. Sepehri-Amin, H., Dirba, I., Tang, X., Ohkubo, T., Schrefl, T., Gutfleisch, O., Hono, K.: Acta
Mater. 175, 276 (2019)
128. Moore, J., Klemm, D., Lindackers, D., Grasemann, S., Träger, R., Eckert, J., Löber, L.,
Scudino, S., Katter, M., Barcza, A., Skokov, K., Gutfleisch, G.: J. Appl. Phys. 114, 043907
(2013)
129. Jacimovic, J., Binda, F., Herrmann, L., Greuter, F., Genta, J., Calvo, M., T. Tomse, Simon, R.:
Adv. Eng. Mater. 19, 1 (2017)
130. Huber, C., Abert, C., Bruckner, F., Groenefeld, M., Muthsam, O., S. Schuschnigg, Sirak, K.,
Thanhoffer, R., Teliban, I., Vogler, C., Windl, R., Suess, D.: Appl. Phys. Lett. 109, 162401
(2016)
131. Huber, C., Abert, C., Bruckner, F., Pfaff, C., Kriwet, J., Groenefeld, M., Teliban, I., Vogler,
C., Suess, D.: J. Appl. Phys. 122, 053904 (2017)
132. Li, L., Tirado, A., Nlebedim, I., Rios, O., Post, B., Kunc, V., Lowden, R., Lara-Curzio, E.,
Fredette, R., Ormerod, J., Lograsso, T., Paranthaman, M.: Sci. Rep. 6, 1 (2016)
133. Paranthaman, M., Shafer, C., Elliot, A., Siddel, D., McGuire, M., Springfield, R., Martin, J.,
Fredette, R., Ormerod, J.: JOM 68, 1978 (2016)
134. Schneider, J., Eckert, D., Müller, K.-H., Handstein, A., Mühlbach, H., Sassik, H., Kirchmayr,
H.R.: Mater. Lett. 9, 201 (1990)
135. Coehoorn, R., De Mooij, D.B., Duchateau, J.P.W.B., Buschow, K.H.J.: Le J. De Physique
Colloques 49, 669 (1988)
136. Kneller, E.F., Hawig, R.: IEEE Trans. Magn. 27, 3588 (1991)
137. Meiklejohn, W.H., Bean, C.P.: Phys. Rev. 105, 904 (1957)
138. Bollero, A., Gutfleisch, O., Müller, K.-H., Schultz, L., Drazic, G.: J. Appl. Phys. 91, 8159
(2002)
139. Hirosawa, S.: Trans. Magnetics Soc. Jpn. 4, 101 (2004)
140. Zhang, P., P.Minxiang, A.N., J.I.Zhiwei, A.O., Qiong, W.U., Hongliang, G.E., Rui, F.U.: J.
Rare Earths 28, 944 (2010)
141. Skomski, R., Coey, J.M.D.: Phys. Rev. B 48, 15812 (1993)
142. Hirosawa, S.: J. Magn. Soc. Jpn. 39, 85 (2015)
143. Sawatzki, S., Heller, R., Mickel, C., Seifert, M., Schultz, L., Neu, V.: J. Appl. Phys. 109,
123922 (2011)
144. Liu, X., He, S., Qiu, J.M., Wang, J.P.: Appl. Phys. Lett. 98, 222507 (2011)
145. Betancourt, I., Davies, H.A.: Mater. Sci. Technol. 26, 5 (2010)
146. Coey, J.M.D., Sun, H.: J. Magn. Magn. Mater. 87, L251 (1990)
147. Coey, J.M.D., Sun, H., Otani, Y., Hurley, D.P.F.: J. Magn. Magn. Mater. 98, 76 (1991)
148. Dirken, M.W., Thiel, R.C., De Jongh, L.J., Jacobs, T.H., Buschow, K.H.J.: J. Less-Common
Met. 155, 339 (1989)
149. Sun, H., Otani, Y., Coey, J.M.D.: J. Magn. Magn. Mater. 104, 1439 (1992)
150. Coey, J.M.D., Smith, P.A.I.: J. Magn. Magn. Mater. 200, 405 (1999)
1430 K.-H. Müller et al.

151. Sun, H., Coey, J.M.D., Otani, Y., Hurley, D.P.F.: J. Phys. Condens. Matter 2, 6465 (1990)
152. Zhong, X.-P., Radwański, R.J., De Boer, F.R., Jacobs, T.H., Buschow, K.H.J.: J. Magn. Magn.
Mater. 86, 333 (1990)
153. Fujii, H., Sun, H., In: K.H.J. Buschow (ed.) Handbook of Magnetic Materials, vol. 9. North-
Holland, Amsterdam (1995), pp. 303–404
154. Kataoka, K., Matsuura, M., Tezuka, N., Sugimoto, S.: Mater. Trans. 56, 1698 (2015)
155. Skomski, R., Müller, K.-H., Wendhausen, P.A.P., Coey, J.M.D.: J. Appl. Phys. 73, 6047 (1993)
156. Li, H.S., Coey, J.M.D.: In: K.H.J. Buschow (ed.) Handbook of Magnetic Materials, vol. 6.
North-Holland, Amsterdam (1991), p. 1
157. Skomski, R.: In: Coey, J.M.D. (ed.) Ch. 4 in Rare-Earth – Iron Permanent Magnets. Oxford
University Press (1996)
158. Machida, K., Adachi, G.: In: (ed.) S.T. Oyama Ch. 10 in The Chemistry of Transition Metal
Carbides and Nitrides. Springer Netherlands (1996)
159. Reller, A.: physica status solidi (RRL)-Rapid Research Letters 5, 309 (2011)
160. Graedel, T.E., Barr, R., Chandler, C., Chase, T., Choi, J., Christoffersen, L., Friedlander, E.,
Henly, C., Jun, C., Nassar, N.T., et al.: Environ. Sci. Technol. 46, 1063 (2012)
161. Gauß, R., Gutfleisch, O.: Magnetische Materialien – Schlüsselkomponenten für neue Energi-
etechnologien, in Rohstoffwirtschaft und gesellschaftliche Entwicklung. Springer (2016)
162. European Commission. Report on critical raw materials for the EU – Report of the Ad hoc
Working Group on defining critical raw materials. Critical raw materials profiles (2014)
163. U.S. Department of Energy. Critical materials strategy (2011)
164. Overschott, K.J.: IEEE Proc A 138, 22 (1991)
165. USGS. Rare earths statistics and information. Accessed on 22 Apr 2016 (2016)
166. Metal-pages. Teddington, www.Metalpages.com. Accessed on 23 May 2016 (2016)
167. Binnemans, K.: In: Proceedings of the 1st European Rare Earth Resources Conference, Milos
(Greece) (2014), pp. 37–46
168. Simon, D., Wuest, H., Hinderberger, S., Koehler, T., Marusczyk, A., S. Sawatzki, Diop, L.,
Skokov, K., Maccari, F., Senyshyn, A., Ehrenberg, H., Gutfleisch, O.: Acta Mater. 172, 131
(2019)
169. Skokov, K., Gutfleisch, O.: Scripta Mater. View Point Set 154, 289 (2018)
170. Poenaru, I., Lixandru, A., Riegg, S., Fayyazi, B., Taubel, A., Güth, K., Gauß, R., Gutfleisch,
O.: J. Magn. Magn. Mater. 478, 198 (2019)
171. Constantinides, S.: Permanent magnets in a changing world market, Magnetics Business and
technology. Accessed on 22 Apr 2016 (2016)
172. Haxel, G.B., Hedrick, J.B., Orris, G.J.: USGS: Rare Earth Elements Critical- Resources for
High Technology (2002)
173. ERECON: Strengthening the European rare earth supply-chain. Challenges and policy
options. A report by the European rare earths competency network (ERECON). Accessed
on 22 Apr 2016 (2016)
174. Gschneidner, K.A., McCallum, R.W., Khan, M., Pathak, A.K., Pecharsky, V.K., Zhou, L.,
Sun, K., Dennis, M.J., Kramer, K.W.: In: Hadjipanayis, G.C., Chen, C.H., Liu, J.P. (eds.)
REPM Proceedings of 23rd International Workshop on RE Magnets and their Applications
(2014), pp. 403–405
175. McKinsey & Company: Lighting the way: Perspectives on the global lighting market.
Accessed on 22 Apr 2016 (2012)
176. Schüler, D., Buchert, M., Liu, R., Dittrich, S., Merz, C.: Study on rare earths and their
recycling. final report for the greens/efa group in the european parliament. Accessed on 22
Apr 2016 (2011)
177. Kingsnorth, D.: The rare earths industry: sustainable or stagnant? 2014 forecast, presentation
at the ERECON Steering Committee in Brussels, June 26, slide 14 (2014)
178. Sprecher, B., Xiao, Y., Walton, A., Speight, J., Harris, R., Kleijn, R., Visser, G., Kramer, G.J.:
Environ. Sci. Technol. 48, 3951 (2014)
179. Prnewswire: Global and China NdFeB industry report, 2015–2018. Accessed on 22 Apr 2016
(2016)
28 Permanent Magnet Materials and Applications 1431

180. Roskill: Rare earths: Global industry, markets and outlook to 2026, 16th edn, p. 174 (2016)
181. Walton, A., Williams, A.: Mater. World 19, 24 (2011)
182. Gutfleisch, O., Güth, K., Woodcock, T.G., Schultz, L.: Adv. Energy Mater. 3, 151 (2012)
183. Jack, K.H.: In: Proceedings of the Royal Society of London A: Mathematical, Physical and
Engineering Sciences, pp. 200–215. The Royal Society (1951), 1093
184. Jack, K.H.: In: Materials Science Forum. Trans Tech Publications (2000), pp. 91–98
185. Ogawa, T., Ogata, Y., Gallage, R., Kobayashi, N., Hayashi, N., Kusano, Y., Yamamoto, S.,
Kohara, K., Doi, M., Takano, M., et al.: Appl. Phys. Express 6, 073007 (2013)
186. Dirba, I., Schwöbel, C., Diop, L., Dürrschnabel, M., Molina-Luna, L., Hofmann, K.,
Komissinskiy, P., Kleebe, H.-J., Gutfleisch, O.: Acta Mater. 123, 214 (2017)
187. Zhang, H., Dirba, I., Helbig, T., Alff, L., Gutfleisch, O.: APL Mater. 4, 116104 (2016)
188. Dirba, I., Komissinskiy, P., Gutfleisch, O., Alff, L.: J. Appl. Phys. 117, 173911 (2015)
189. Widenmeyer, M., Niewa, R., Hansen, T.C., Kohlmann, H.: Zeitschrift für anorganische und
allgemeine Chemie 639, 285 (2013)
190. Poirier, E., Pinkerton, F.E., Kubic, R., Mishra, R.K., Bordeaux, N., Mubarok, A., Lewis, L.H.,
Goldstein, J.I., Skomski, R., Barmak, K.: J. Appl. Phys. 117, 17E318 (2015)
191. Lewis, L., Mubarok, A., Poirier, E., Bordeaux, N., Manchanda, P., Kashyap, A., Skomski,
R., Goldstein, J., Pinkerton, F.E., Mishra, R.K., et al.: J. Phys. Condens. Matter 26, 064213
(2014)
192. Kota, Y., Sakuma, A.: J. Phys. Soc. Jpn. 83, 034715 (2014)
193. Gong, M., Kirkeminde, A., Wuttig, M., Ren, S.: Nano Lett. 14, 6493 (2014)
194. Coey, J.M.D.: J. Phys. Condens. Matter 26, 064211 (2014)
195. Kuzmin, M.D., Skokov, K.P., Jian, H., Radulov, I., Gutfleisch, O.: J. Phys. Condens. Matter
26, 064205 (2014)
196. Edström, A., Werwiński, M., Iuşan, D., Rusz, J., Eriksson, O., Skokov, K.P., Radulov, I.A.,
Ener, S., Kuz’min, M.D., Hong, J., et al.: Phys. Rev. B 92, 174413 (2015)
197. Kōno, H.: J. Phys. Soc. Jpn. 13, 1444 (1958)
198. Mix, T., Bittner, F., Müller, K.-H., Schultz, L., Woodcock, T.G.: Acta Mater. 128, 160
(2017)
199. Braun, P.B., Goedkoop, J.A.: Acta Crystallogr. 16, 737 (1963)
200. Odahara, H., Tomiyoshi, S., Shinohara, T.: Phys. B Condens. Matter 237, 568 (1997)
201. Pasko, A., Mazaleyrat, F., Varga, L.K., Stamenov, P.S., Coey, J.M.D.: IEEE Trans. Magn. 50,
1 (2014)
202. Wiezorek, J.M., Kulovits, A.K., Yanar, C., Soffa, W.A.: Metall. Mater. Trans. A 42, 594
(2011)
203. Bittner, F., Schultz, L., Woodcock, T.G.: Acta Mater. 101, 48 (2015)
204. Sakamoto, Y., Kojima, S., Kojima, K., Ohtani, T., Kubo, T.: J. Appl. Phys. 50, 2355 (1979)
205. Ohtani, T., Kato, N., Kojima, S., Kojima, K., Sakamoto, Y., Konno, I., Tsukahara, M., Kubo,
T.: IEEE Trans. Magn. 13, 1328 (1977)
206. Yamaguchi, A., Tanaka, Y., Yanagimoto, K., Sakaguchi, I., Kato, N.: Bull. Jpn. Inst. Met. 28,
422 (1989)
207. Tanaka, Y., Aikawa, Y.: J. Jpn. Soc. Powder Powder Metall.(Jpn.) 47, 15 (2000)
208. Zeng, Q., Baker, I., Cui, J.B., Yan, Z.C.: J. Magn. Magn. Mater. 308, 214 (2007)
209. Jian, H., Skokov, K.P., Gutfleisch, O.: J. Alloys Compd. 622, 524 (2015)
210. El-Gendy, A.A., Hadjipanayis, G.: J. Phys. D. Appl. Phys. 48, 125001 (2015)
211. El-Gendy, A.A., Hadjipanayis, G.C.: J. Phys. Chem. C 119, 8898 (2015)
212. Ener, S., Skokov, K.P., Karpenkov, D.Y., Kuz’min, M.D., Gutfleisch, O.: J. Magn. Magn.
Mater. 382, 265 (2015)
213. Mix, T., Müller, K.-H., Schultz, L., Woodcock, T.G.: J. Magn. Magn. Mater. 391, 89 (2015)
214. Herbst, J.F., Meyer, M.S., Pinkerton, F.E.: J. Appl. Phys. 111, 07A718 (2012)
215. Hussain, M., Zhao, L.Z., Zhang, C., Jiao, D.L., Zhong, X.C., Liu, Z.W.: Phys. B Condens.
Matter 483, 69 (2016)
216. Chen, D.X., Brug, J.A., Goldfarb, R.B.: IEEE Trans. Magn. 27, 3601 (1991)
217. Coey, J.M.D.: J. Magn. Magn. Mater. 248, 441 (2002)
1432 K.-H. Müller et al.

218. Hasegawa, M., Uchida, K., Nozawa, Y., Endoh, M., Tanigawa, S., Sankar, S.G., Tokunaga,
M.: J. Magn. Magn. Mater. 124, 325 (1993)
219. Kirchner, A., Hinz, D., Panchanathan, V., Gutfleisch, O., Müller, K.-H., Schultz, L.: IEEE
Trans. Magn. 36, 3288 (2000)
220. Sawatzki, S., Dirba, I., Schultz, L., Gutfleisch, O.: J. Appl. Phys. 114, 133902 (2013)
221. Takeshita, T., Morimoto, K.: J. Appl. Phys. 79, 5040 (1996)
222. Sawatzki, S.: Der Korngrenzendiffusionsprozess in nanokristallinen Nd-Fe-B-
Permanentmagneten (The grain boundary diffusion process in nanocrystalline Nd-Fe-B
permanent magnets). http://tuprints.ulb.tu-darmstadt.de/id/eprint/5221 (2015)
223. Abele, M.: Structures of Permanent Magnets. Wiley, New York (1993)
224. Leupold, H.: In: Coey J.M.D. (ed.) Ch. 8: Rare-earth iron permanent magnets. Oxford
University Press (1996)
225. Zijlstra, H.: Phillips J. Res. 40, 259 (1985)
226. Bjørk, R., Bahl, C.R.H., Smith, A., Pryds, N.: J. Magn. Magn. Mater. 322, 3664 (2010)
227. Bjørk, R., Bahl, C.R.H, Nielsen, K.K.: Int. J. Refrig. 63, 48 (2016)
228. Bloch, F., Cugat, O., Meunier, G., Toussaint, J.C.: IEEE Trans. Magn. 2465–2468 (1998)
229. CERN Courier: European Synchrotron Radiation Facility ESRF, vol.43, No3, p.7, (2002)
230. Cugat, O., Hansson, P., Coey, J.M.D.: IEEE Trans. Magn. 30, 4602 (1994)
231. Widmer, J.D., Martin, R., Kimiabeigi, M.: Sustain. Mater. Technol. 3, 7 (2015)
232. Borisavljevic, A.: Limits, modeling and design of high-speed permanent magnet machines.
Springer Science & Business Media (2012)
233. Binder, A., Schneider, T.: In: Electrical Machines and Power Electronics, 2007. ACEMP’07.
International Aegean Conference on IEEE, pp. 9–16 (2007)
234. Bang, D., Polinder, H., Shrestha, G., Ferreira, J.A.: In: European Wind Energy Conference &
Exhibition, Belgium, pp. 1–11 (2008)

Karl-Hartmut Müller received his PhD from Dresden Univer-


sity of Technology (TU Dresden) in 1968. The main fields of his
investigations at IFW Dresden are magnetism and magnetic mate-
rials as well as superconductivity and superconducting materials.
He also worked with other systems, e.g., fullerides and metallic
oxides (e.g., manganites). Furthermore, he is interested in ther-
modynamics and the description of non-equilibrium processes.
28 Permanent Magnet Materials and Applications 1433

Simon Sawatzki received his PhD from the Technical Univer-


sity Darmstadt in 2015, where he was active in the field of
processing and characterization of Nd-Fe-B permanent magnets,
especially the hot deformation and grain boundary diffusion pro-
cess. From 2018 on, he works as a researcher at Vacuumschmelze
GmbH & Co. KG with focus on magnetic cores and inductive
components for filter applications.

Oliver Gutfleisch is a full professor of Functional Materials


at Technical University Darmstadt. His interests span from new
permanent magnets for power applications to energy-efficient
magnetic cooling, ferromagnetic shape memory alloys, and mag-
netic nanoparticles for biomedical applications, with particular
emphasis on tailoring structural and chemical properties on the
nanoscale. Resource efficiency and recycling of strategic metals
are also in the focus of his work.

Roland Gauß received his PhD from the University of Tübingen


in 2008. His research interests are related to metallurgy, the
resource criticality and life cycles of materials, and innovation.
He worked in the fields of Nd-Fe-B processing and recycling, the
extractive metallurgy of copper, and geochemical fingerprinting.
He joined EIT RawMaterials from Fraunhofer IWKS.
Soft Magnetic Materials and Applications
29
Frédéric Mazaleyrat

Contents
Losses in Soft Magnets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1436
Hysteresis Loss . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1438
Low-Frequency Losses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1438
High-Frequency Losses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1449
Iron and Low-Carbon Steels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1452
Electrical Steels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1452
Composition, Processing, and Texture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1452
Non-oriented Silicon Steel Sheets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1455
Grain-Oriented Silicon Steel Sheets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1457
Trends in Fe-Si Products . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1460
Iron-Cobalt Alloys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1461
Equiatomic Fe-Co Alloy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1461
Low-Cobalt Alloys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1463
Iron-Nickel Alloys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1463
Ni-Rich Alloys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1463
Fe-Rich Alloys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1465
Thermal Alloys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1466
Amorphous and Nanocrystalline Alloys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1467
Iron-Based Amorphous Alloys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1468
Cobalt-Based Amorphous Alloys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1470
Nanocrystalline Alloys . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1470
Soft Ferrites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1472
Spinel Ferrites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1472
Synthesis of Sintered Ferrites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1473
Mn-Zn Ferrites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1474
Other Soft Ferrites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1478

F. Mazaleyrat ()
SATIE, CNRS, École Normale Supérieure Paris-Saclay, Gif-sur-Yvette, France
e-mail: mazaleyr@ens-paris-saclay.fr

© Springer Nature Switzerland AG 2021 1435


J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_31
1436 F. Mazaleyrat

Effect of a Gap on Magnetic Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1481


Core with Lumped Gap . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1481
Cores with Spread Gap . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1481
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1486

Abstract

A great variety of soft magnetic material has been developed with the aim
of enhancing the magnetization, increasing the permeability, controlling the
hysteresis loop shape, or raising the working frequency while in all cases
decreasing the magnetic losses. Besides the physical considerations, material
shaping and cost are important parameters. It is thus not surprising to see
that despite natural selection, hundreds of materials of very different natures
(metallic, ceramic, glass) and properties remain on the market. Choosing the
right material for a specific application is often a tricky compromise and, too
often, a question of industrial culture. This chapter aims at helping the user to
make the best choice possible and gives some keys to improving device design.
The first part is devoted to magnetic losses, giving the most used models in
the low-frequency nonlinear regime and the high-frequency linear regime. The
following section describes the different families of materials: low-alloyed steels,
iron-cobalt alloys, iron-nickel alloys, amorphous and nanocrystalline alloys, and
soft ferrites. In each section, the fabrication process is described, and tables of
typical properties of most common materials are given. Finally, there is a table
of applications and matching materials.

Losses in Soft Magnets

From his work on thermodynamics and magnetism, Lord Kelvin proposed that any
change in magnetization should be accompanied by dynamic losses and heat release
or absorption – an effect that was later discovered and named magnetocaloric effect
by Pierre Weiss. Early work of Cazin in this direction has shown that change in
magnetization at constant temperature always produces heat [16], but due to the
geometry of the system and the frequency of magnetization switching, this was due
essentially to eddy currents. In 1881 Warburg proved that this heating was not only
due to dynamic effects [38] but to some kind of friction too, which was confirmed
independently the same year by Ewing [18] who developed the first B(H ) loop
tracer. He showed that the loop area is identical to heat release and named the
phenomenon hysteresis. Immediately, it was recognized that the hysteresis depends
on both flux density and frequency and that, below a certain frequency, the area and
shape of the loop depends no more on the frequency. This floor value of the losses is
known as hysteresis loss. The concept of loss separation was proposed by Braisford
[15] as the sum of the hysteresis loss, Phy , and eddy current loss, Pec . The difference
29 Soft Magnetic Materials and Applications 1437

750

Bpeak=1.5 T

600

excess loss
450
wtot=(Jm-3)

eddy current loss


300

150
hysteresis loss

0
0 50 100 150 200 250
Frequency (Hz)

Fig. 1 Separation of losses in a 2.4 W/kg grade silicon steel. Lines correspond to the model and +
to the experiment

between measured iron loss PF e and these two components was called anomalous
loss Pan = PF e − Phy − Pec (see Fig. 1). The physical origin of what we rather
called excess loss Pex will be found long time later and properly embedded in a
reliable theory only in 1988 (see section “Excess Loss ”). In general, losses have
to be expressed in specific units, either power or energy per unit volume or mass
depending on customs of different communities. Eventually in linear regime, losses
can be expressed as a function of the loss angle δ = π2 − phase lag between voltage
and current. Usual units are:

– Electrical machines W · kg−1


– Power electronics W · m−3 or mW · cm−3
– Modelling J · m−3
– Electronics tan δ

The dependence of losses upon magnetization and frequency is rather complex


and can hardly be modeled by explicit equations; however, in a first approximation,
losses can be fitted using the separation principle and Steinmetz exponents in the
following forms:
1438 F. Mazaleyrat

Volume energy loss Volume power loss Mass power loss


Expressed in Jm−3 Expressed in Wm−3 Expressed in Wkg−1
Wh = k0 B 2 Ph = k 0 B 2 f ph = k0 B 2 f/ρ
Wcl = k1 B 2 f Pcl = k1 B 2 f 2 pcl = k1 B 2 f 2 /ρ
W ex = k2 B m f n Pex = k2 B m f n+1 pex = k2 B m f n+1 /ρ

It must be however emphasized that only the classical term can be determined
a priori. The other terms are determined from experiments. As it will be shown in
section “Excess Loss ”, the exponents of the excess part are n = 1/2 and m = 3/2
in many cases, but this is not a general rule. In a limited range of induction and
frequency, it is always possible to find an expression of losses like PF e = kB η f ξ ,
but this is only a fitting definitively free of physical sense.

Hysteresis Loss

Hysteresis loss is the most difficult loss term to model. At the end of the nineteenth
century, Rayleigh and Steinmetz proposed empirical models for hysteresis loss, and
we must say that more than one century later, there is no better simple model.
Since the 1950s, the Preisach model has been developed and widely studied. It’s
based on the decomposition of the hysteresis loop in elementary loops (hysteron)
defined by a coercive and a shift field. To compute the Preisach distribution , it
is necessary to record a great number of minor loops (or reversal curves) in order
to sweep uniformly the whole area of the major loop. The distribution allows to
compute any trajectory in the M(H ) plane, but, finally, if it is useful for researchers
to study the magnetization mechanisms in soft or even hard materials, it remains
usually too complicated for implementation in industrial software. The recovery
of the distribution is not straightforward and would need developments out of the
scope of this chapter; for more detail, the reader is invited to refer to the monumental
“Science of Hysteresis” [14].

Low-Frequency Losses

Steinmetz Model
The Steimetz model for hysteresis loss has actually no physical basis. It was
discovered purely empirically in 1905 and written in the form kB η . Plotting
hysteresis losses vs induction in log-log scale allows to deduce the coefficient
and the exponent, but the exact value of the exponent η depends somehow on the
considered induction window as it is seen in Fig. 2. As observed by Steinmetz, a
value close to 1.6 is generally observed. This value can’t be easily connected with
the result of Rayleigh because the latter is expressed as a function of H and because
the permeability is not constant. At higher induction, the exponent increases over
2 and then drops to zero in the reversibility region of the major loop (close to
29 Soft Magnetic Materials and Applications 1439

Fig. 2 Fitting of Steinmetz exponent in a non-oriented Fe-Si steel sheet

saturation) [10]. In average in a large range of induction, an exponent 2 is usually


a good approximation, but the error can be up to 100%. However, in some cases,
hysteresis loss can be a minor component of the loss.

Rayleigh Loops
In the small signal regime, i.e., when B  μ0 MS , Rayleigh observed linear and
quadratic behaviors of permeability and initial magnetization curve, respectively:

μ = μi + νH and B = μi H + νH 2 (1)

where μi is the initial permeability and η is an empirical constant. Starting from a


nonzero field Hm , the upper or lower branch of the hysteresis loops can be described
from this law in the form:

1
B = (μi + νHm )H ± ν(Hm2 − H 2 ) (2)
2

from which the area of the loop can be deduced:


1440 F. Mazaleyrat

4 4
Wh = νH 3 or Ph = νHm3 f (3)
3 m 3

To write the loss dependence on B, one has to solve the second-order equation
(1) which single physical solution can be introduced in (3):
 
μ3 4νB
Wh = i3 1+ 2 −1 (4)
6ν μi

If μ2i  4νB:

4 μ3i 3/2
Wh = B (5)
3 ν

we end up with a Steinmetz exponent of 1.5, which is very close to the experimental
one.
This law is pretty much simple, but its validity is limited to B  JS /10. Such
low induction in practice is used only at high frequency, i.e., several tens or hundreds
of kHz. Under these conditions, hysteresis losses are negligible compared to eddy
currents in metals, so this law is more useful in practice for ferrites. Experience
shows that the Rayleigh law is well obeyed in ferrites up to 25 mT and over 100 kHz
in highly resistive NiZn-type ferrites (see, e.g., [1]).

Eddy Currents
When any conductive material is submitted to AC magnetic field, eddy currents are
created inside in such a way they produce a counter field. This contribution to losses
will be added to the hysteresis (DC) component of the losses. Its nature is purely
Maxwellian and refers to an isotropic and homogeneous material at macroscopic
scale, which is not realistic for ferromagnetic materials because of the domain
structure. However, in certain conditions, i.e., if domains are small compared to
dimensions of the sample and isotropically distributed or if the domain walls move
little compared to domain width (Rayleigh regime), eddy currents are as if the
material is magnetically homogeneous.

Classical Losses
This term refers to eddy current loss in the approximation of a weak skin effect, i.e.,
when the magnetic field produced by eddy currents in reaction to the exciting field,
Hex , is very small. In other worlds, the inductive reaction has no significant effect
on the flux density B which is considered uniform in space. From Maxwell-Faraday
equation

∂B
∇ ×E=− (6)
∂t
29 Soft Magnetic Materials and Applications 1441

w j B

j z
d y
x
L
Fig. 3 Geometry of the problem for a thin sheet; eddy currents are everywhere perpendicular to
B and parallel to the surface

and looking at the geometry of the problem (Fig. 3) with L, w  d, one can see that
(i) the flux density has only one component along x depending only on time and (ii)
the current density, j, has a single y component independent on x and y, so that the
electrical field has the same symmetry through Ohm’s law j = σ E:

∂Ey ∂Bx ∂Bx


= ⇒ Ey (t, z) = z+K (7)
∂z ∂t ∂t

Because of the symmetry, Ey = 0 in the middle – where we advantageously locate


the origin – K = 0. The power dissipated in the volume can be obtained from the
integral on the half thickness:

  2  d/2
∂Bx
Pcl (t) = σ Ey2 dV = 2Lwσ z2 dz (8)
V ∂t 0

and dividing by the volume dwL, the instantaneous power per unit volume is
derived:
 2
1 ∂Bx
Pcl (t) = σ d2 (9)
12 ∂t

Getting the mean power density requires to define the function Bx (t). Usually
in low-frequency applications such as AC machines and transformers, as well as in
standard characterization conditions, the induction is sinusoidal. The average of a
squared sinus being 12 , we end up with the well-known formula, where Bp denotes
the peak value of the induction:

π2 2 2 2
Pclsin = σf Bp d (10)
6

By opposition, in power electronics, the voltage has often a square wave form, so
B(t) is triangular. In this case, the time derivative of B is constant over a half period
so its value is 4f Bp for an induction swing of 2Bp and one gets:
1442 F. Mazaleyrat

4 2 2 2
Pcltri = σf Bp d (11)
3

It is easy to see that the classical losses are 23% lower for triangle waveform
compared to sinus for the same Bp .

Skin Effect
When the low-frequency condition is not satisfied, interaction between electrical
and magnetic fields must be introduced, and Maxwell-Ampère equation can be
combined with Ohm’s law as ∇ × H = j = σ E. Further, the permeability is
introduced:

∂B ∂H
∇ × (∇ × H) = ∇ × j = σ ∇ × E = −σ = −σ μ (12)
∂t ∂t

Using the property ∇ × (∇ × H) = ΔH − ∇(∇ · H) = ΔH, one finally gets the


diffusion equation:

∂H
ΔH = −σ μ (13)
∂t

In the geometrical conditions of Fig. 3,

∂ 2 Hx (z, t) ∂Hx (z, t)


2
= σμ (14)
∂z ∂t

In sinusoidal regime, this differential equation can be rewritten in complex form:

∂ 2 h(z)
= iσ μωhx (z) (15)
∂z2

where the peak value of the field in the direction x is 2h. A solution of this
equation can be found in the form of either exponential or hyperbolic functions.
In this case, because of the symmetry, the surface field is equal to the applied
field h(d/2) = h(−d/2) = ha so the hyperbolic solution has the simple form
hx (z) = A cosh(kz) where the complex number k is define as k2 = iσ μω. It is
usually convenient to write k = (1 + i)/δ, where the quantity δ, homogeneous with
a length, is called the penetration or skin depth:

2
δ= (16)
σ μω

so a hyperbolic solution is easily found taking into account the boundary conditions:
29 Soft Magnetic Materials and Applications 1443

cosh(kz)
h(z) = ha (17)
cosh(d/2)

The RMS value of the field is obtained by calculating the square root of the
conjugate product:

cosh (2z/δ ) + cos (2z/δ )
h(z) = ha (18)
sinh (d/δ ) + sin (d/δ )

With the present symmetry of the problem, the RMS value of the y single
component of the current density is written from the Maxwell-Ampère equation
as
√ 
dh 2 cosh (2z/δ ) − cos (2z/δ )
j (z) = = ha (19)
dz δ cosh (d/δ ) + cos (d/δ )

The volumetric Joule losses produced by eddy currents in the sample are obtained
by integration over the thickness divided by the volume

1 d/2  
Pec = jy (z)2 dz (20)
σe −d/2

2 sinh (d/δ ) − sin (d/δ )


Pec = Ha2 , (21)
σ eδ cosh (d/δ ) + cos (d/δ )

where Ha is the peak value of the applied (external) field.


However, in many cases – electrical machines, transformers, and standard
measurements – not the excitation field is the driving force but the induction.
Considering a constant permeability in the whole material, the average induction
over the cross section is written knowing the value of B at the surface (x, y, ±e/2):
 d/2

1 cosh (kz) tanh kd/2


B(z) = Bsurf dz = Bsurf (22)
d −d/2 cosh (kd/2) (kd/2)

The quantity Hex in (21) is substituted by B(z)/μ0 μR , and calculating the


conjugate quantity product of the hyperbolic term yields the following reduction:

πf d B2 sinh dδ − sin dδ


Pcl = ·

(23)
2δ μ0 μR cosh dδ − cos dδ

It must be emphasized that this formula is valid only when the permeability is
constant. This condition is a strong restriction when δ  d because the field at
the surface is much stronger than in the center (see Fig. 4). As a consequence, even
1444 F. Mazaleyrat

2 2

1.8 1.8

1.6 1.6

1.4 1.4

1.2 1.2
H/H ex

B/ B
1 1

0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2

0 0
-0.5 0 0.5 -0.5 0 0.5
z/t z/t

Fig. 4 Profile of the field H across the thickness reduced to excitation field and profile of B
reduced to the space averaged induction for δ/d =1, 0.5, 0.1. B is calculated in the linear case;
clearly, this model is not adapted for small δ/d when B is close to JS

if B  JS , the sample may be saturated close to the surface. In this case, there
is no analytical solution; specific and time-consuming algorithms must be used (see
[17] for more detail.)

Excess Loss

The Pry and Bean Model


The well-known Pry and Bean model was developed to study the eddy current losses
generated by the movement of domain walls in a periodic array of infinite slab
domains. Let’s call 2L the width of a domain, d the thickness of the slabs, and x the
instantaneous position of the wall with respect to the zero net magnetization of the
array (x = 0 for B = 0). The relation between the instantaneous magnetization
and the position of the wall is thus J (t) = L1 x(t)JS . Taking into account the
boundary conditions, the solution can be express in terms of series expansion of
hyperbolic terms. The resolution is rather complicated, but details can be found in
the classical paper [29]. For convenience, we adopted the following form of the
time-dependent losses per unit volume [11]:
 2    
σ ∂φ 4 1 L+x L−x
PP B (t) = coth nπ + coth nπ
2Ld ∂t π4 n3 d d
odd n
(24)
29 Soft Magnetic Materials and Applications 1445

If the domain are very thin (2L  d) coth(x) 1/x and one can find the
classical result of (9). On the overside, if 2L  d, the coth terms are close to unity
and (24) is reduced to:

 2
σG ∂φ
PP B (t) = (25)
2Ld ∂t


by writing 4
π4
1
n3
= G = 0.1356. In grain-oriented Fe-Si sheets, the latter
odd n
condition is rather justified since 2L ≈ 1 mm; d ≈ 0.24 mm and very well justified
for longitudinal field-annealed amorphous ribbons 2L ≈ 5 mm; d ≈ 0.02 mm.

Bertotti’s Model
The basis of Bertotti’s model (see [12] and [11] for more detail) is the application
of Pry and Bean model to a system composed of slab domain regions randomly
oriented in space. From statistical arguments, Bertotti has shown that in this case the
total dynamic loss is equivalent to the sum of classical eddy currents superimposed
with eddy currents generated by periodic slab domains parallel to the field Pdyn (t) =
Pcl (t) + PP B (t). This validates theoretically the concept of loss separation PF e =
Phy + Pcl + Pex , where the excess loss is the time average of (25):
 2   2
σG ∂φ σG ∂φ
Pex = = (1 + α) (26)
2Ld ∂t 2Ld ∂t

where α is a form factor depending only on the flux waveform. It is now interesting
to introduce the notion of excess field which is the supplementary field necessary
to obtain the desired change in magnetization due to the dynamical movement of
domain walls:

∂J
Pex = Hex (27)
∂t

It is now easy to substitute the flux by the magnetization J = φ/2Ld in this


expression and to equate with (26) and to get:

∂J
Hex = 2Ldσ G(1 + α) (28)
∂t

Let us now introduce the width of the sheet, ; the area of the cross section
perpendicular to the field, S = d; and the number of domain walls in this section,
nw = /2L. Now we have:

S ∂J σ GS ∂J
Hex = σ G(1 + α) · = · (29)
nw ∂t n ∂t
1446 F. Mazaleyrat

nw
where we substituted n = 1+α to simplify. Thus, n is no more the real number of
domain wall but an effective number depending on the waveform. This is actually
not a problem since n, also called the number of magnetic objects, will be identified
from loss measurements using a given waveform. However, it is reasonable to think
that the number of active magnetic object will depend on the excess field (larger
excess field will activate more magnetic objects), so at first order one can write:

Hex
n = n0 + (30)
V0

where n0 is the number of active objects at zero field and V0 the field necessary to
activate one more object over the total number of magnetic objects available N0 .
Physically speaking, a magnetic object can be a domain or a set of domain moving
coherently, in other words, a crystallographic grain. Introducing (30) in (29) leads
to the second-order equation:

2
Hex ∂J
n0 Hex + n0 = σ GS (31)
V0 ∂t

which admits one physical solution over the two mathematical ones

−n0 V0 + n20 V02 + 4V0 σ GS ∂J
∂t
Hex = (32)
2

The excess loss can now be expressed from (26) after factorization:
 
n0 V0 4σ GS ∂J ∂J
Pex = 1+ 2 · −1 (33)
2 n0 V0 ∂t ∂t

Deducing the parameters of the model from measurement is not straightforward,


but there is an abundant literature describing the method. By chance in many cases,
particularly small grain non-oriented materials, both n0 and V0 are relatively small,
so (33) is reduced to

   3
∂J 2
Pex = σ V0 GS (34)
∂t

So obviously in this case the parameter V0 is the most important from physical
point of view. Indeed, in small grain materials, N0 can be understood as the number
of grains of average diameter D in the cross section and

2Hhy
V0 (35)
N0
29 Soft Magnetic Materials and Applications 1447

where Hhy , the hysteresis field, is defined from Phy the same way the excess
field is defined (see 27). So in principle this parameter is directly related to the
microstructure and the time dependents of the magnetization.
In the case of triangular flux waveform, the time derivative of the magnetization
can be simply expressed on a half period as a function of the frequency and the peak
magnetization ∂J ∂t = 4Jp f , so (33) can be written in terms of power or energy per
unit volume:
 3 3
Pex = 8 σ V0 GSJp2 f 2 (36)
 3 1
Wex = 8 σ V0 GSJp2 f 2 (37)

so we end up with the famous square root frequency dependence of the excess losses
observed experimentally by Brailsford [15]. The classical result of Bertotti (37)
must be considered with caution:

– The approximation n0 = 0 is not always verified, so (33) must be used.


– V0 is an average over a cycle, so in principle, it increases appreciably with Jp
since the number of magnetic object available for switching goes to zero close to
saturation.
– In the case of sine waveform, the factor 8 in (37) should be replaced by 8.76 [8].
– If both n0 and N0 are large, (33) can be approximated by Pex = 4σnGS 0
Jp2 f 2 .
– The decomposition is not possible for significant skin effect; in particular, excess
loss can appear negative at high induction due to the strong induction gradient in
the lamination.

Typical example of square root dependence of the excess losses in non-oriented


silicon steels is given in the left box of Fig. 5 The difference between experimental

Fig. 5 Plot of losses as a function of square root frequency in non-oriented Fe-Si sheets used for
determination of Bertotti’s model parameters (left). V0 parameter evolution with magnetization
(right) after [8]
1448 F. Mazaleyrat

points and the line at low field is due to the fact n0 cannot be assumed exactly zero.
The right box shows that V0 depends on the peak induction and is not a constant as
it is expected from (30). This however makes physical sense as explained above and
can be related to the Preisach distribution. This doesn’t invalidate Bertotti’s model,
but one has to make a choice between accuracy (which needs determination of V0
dependence on (Jp )) and simplicity (using an average V0 ).
For accurate modeling of high-frequency losses with skin effect, more compli-
cated techniques must be used as in reference [10] based on dynamic Preisach model
[13].

Rotational Losses
Rotational losses are important in classical AC machines where the induction has
usually a constant magnitude and a constant angular speed. It is however a simplified
view because locally the induction locus can be rather ellipsoidal than circular.
Also in three-phase transformers, at the junction between columns and yokes, the
induction has a direction depending on time. In the frame of this chapter, only the
case of circular locus will be discussed.

Rotational Eddy Current Loss


Here again, the only part of the losses we can model easily is the eddy current part.
In the case of a circular locus, B(t) = Bp and θ = Ωt. The Ferraris theorem shows
that any rotating field can be decomposed in two fixed direction time-dependent
perpendicular components:

Bx = B cos Ωt and By = B sin Ωt (38)

It follows immediately that

PclR = 2Pcl (39)

Rotational Hysteresis Loss


The most striking figure of rotational hysteresis loss is the fact they first increase
with B at double speed compared to the hysteresis loss measured in a fixed direction,
goes through a maximum around the inflexion point of the magnetization curve,
and drops to zero close to the saturation. This phenomenon discovered by Baily in
1896 seemed first to be inconsistent but soon became a strong argument in favor of
Weber’s molecular theory of magnetism saying that Ampère’s molecular currents
were permanent and that only the direction of the moments changes. Actually,
today’s interpretation is not so far: as the field is enough to saturate locally all
grains (in other words to suppress domain walls), the magnetization is uniform and
turns toward the field lossless. Only the unavoidable eddy currents remain, because
experience can hardly be done in real DC conditions. This is perfectly illustrated
by the measurements shown in Fig. 6 where we see that the calculated classical loss
curve intercepts the measured total loss curve.
29 Soft Magnetic Materials and Applications 1449

Fig. 6 Rotational losses measured at 50 Hz and separation of losses in NOSS [7]

Because evaluation of excess loss is time-consuming and because they are


usually small compared to other components, they are often neglected.

High-Frequency Losses

From the previous sections, it can be concluded that magnetic power losses
approximately go with the square of B and an exponent of f between 1 and 2. As
a consequence, reasonable losses at high frequency can be obtained only keeping a
reasonable value of the B × f product, which means a very low value of B. In this
case, the material works in the linear regime, and its properties are advantageously
represented in the form a complex permeability. An inductance can be modeled
either by a parallel or series circuit, the latter being the most common:
   2 
R N μ0 μR A R
Z = R + j Lω = j ω L − i = jω −i (40)
ω  ω

where N is the number of turns of the coil and A and  are the effective cross-
sectional area and length of the magnetic circuit. From this equation, it is easy to
identify the real part of the permeability and the imaginary part:

(r − R0 ) 
μ = μR − j = μ − j μ (41)
N 2 μ0 S

where R0 is the ohmic resistance of the coil and r is the series resistance standing
for the iron loss. The loss factor is usually defined as:
1450 F. Mazaleyrat

tan δ = μ /μ (42)

High-Frequency Losses in Metals


It is obvious that in the case of high frequency, losses go with the higher frequency
exponent term of the three components, namely, eddy current term in the case
of magnetic conductors. In general, because of eddy currents, metallic magnetic
materials are used in the form of thin ribbons or layers from several μm down to
several nm, so this still corresponds to the usual thin foil approximation studied
previously (see Fig. 3).
If the excitation field is very small, the permeability can be considered more or
less locally independent of Hex and f , so the complex apparent permeability is
simply the ratio between the average induction as defined in (23) and the field:

tanh(kd/2)
μ = μS (43)
(kd/2)

Getting an analytical expression of the real and imaginary parts of (43) is not
straightforward, but any mathematical software will easily compute μ and μ
knowing the main physical parameters: low-frequency permeability, conductivity,
and thickness. The permeability spectra have two main features:

– The static permeability μS = limf →0 μ (f )


– The relaxation frequency max(μ (f )) = μ (fR )

and


fR = (44)
π μ0 μS d 2

so it is easily seen that the product μS fR is a factor of merit independent of the


permeability and proportional to ρ/d 2 .

High-Frequency Losses in Insulators


The case of insulators is much more complicated than that of metals. Indeed, the
exact mechanism producing magnetic losses is not clear and difficult to model. As
a consequence, the first idea is to adopt the Debye model where the mechanism
of losses is analogous to dry friction. In other words, the losses in terms of power
are linearly dependent on frequency. Because in real materials there is always a
distribution of local properties (the local permeability is strongly dependent on the
grain size), there is neither a perfect Debye behavior, so the following model is often
used:
μS
μ=
α (45)
1 + (j f/fR )β
29 Soft Magnetic Materials and Applications 1451

where the models are called:

– Debye if α = β = 1,
– Davidson-Cole if α < 1 and β = 1
– Cole-Cole if β < 1 and α = 1

This model when α and β = 1 is sometimes referred as non-integer derivative,


but the reader must keep in mind that it’s only a mathematical fitting. Some materials
also show two distinct loss mechanisms (usually attributed to domain wall relaxation
and spin resonance) which can be decomposed in two terms [36], but here again it’s
often mainly a fitting (see Fig. 7). Also, ferrites sometimes show strong asymmetric
resonance behavior which cannot be properly modeled [31].
Interestingly, for a given ferrite composition, if the static permeability can be
tuned by controlling grain size and porosity, the merit factor μS fR is constant, and
strikingly for Ni-Zn properly processed it remains between 3 and 5 GHz. Moreover,
if the frequency is far above fR , (45) for α = β ≈ 1 becomes

μS fR
μ→ (46)
f

This asymptotic behavior is known as Snoeck’s limit . It is applicable only for


insulators (Ni-Zn, Mn-Mg ferrites, garnets); for Mn-Zn, additional eddy current loss
has to be taken into account like in metals.

Fig. 7 Complex permeability spectra of different soft magnetic materials. Left, nanocrystalline
wound core: points are experimental, and line is computed using (43) using parameters measured
independently (t=25μm, ρ = 120 · 10−6 m). Right, ferrite showing double relaxation behavior
typical of material having bimodal grain size distribution: points are experimental, and lines are
fitted using the sum two terms with parameters α1 = 0.5, β1 = 0.65, fR1 = 30 kHz, α2 = 1,
β2 = 0.7, and fR2 = 30 MHz (45)
1452 F. Mazaleyrat

Iron and Low-Carbon Steels

Iron was the only soft material used until the very beginning of the twentieth century.
For industrial applications, such as electrical motors and transformers, iron has
usually two functions: magnetic and mechanical. From magnetic point of view,
iron has a high induction and high permeability, but its conductivity is too high for
efficient AC applications. For this reason, iron is today used only in DC applications
such as pole pieces for lab electromagnets. The mechanical function of iron is
very important because the magnetic core carries the coils in transformers and is
subject to magnetic forces in machines or electromagnets. As a consequence, low-
carbon steel is usually preferred for DC application, whereas it can’t be used in AC
application because of the enhancement of the coercivity due to interstitial C atom in
the bcc lattice of iron. The main application of low-carbon steels as magnetic parts
is: rotors of current excited synchronous machines and stators of DC machines.
In power station turbogenerators and claw pole automobile generators, the rotor is
made by forging. In the case of DC machines, the stator can be made by stamping
of thick sheet for several hundred watt machines and casting for bigger machines.

Electrical Steels

Composition, Processing, and Texture

Composition
In the beginning of the industrial era, metallurgists tried to reduce the carbon
residue in iron to enhance its magnetic softness at the price of poor mechanical
properties. In 1896, Hadfield fortuitously produced steel with a small percentage
of silicon and observed strong enhancement of its mechanical hardness. By 1900,
he measured its magnetic properties, showing that silicon not only improves iron’s
mechanical properties but also resistivity without prejudice to magnetic properties.
A couple of years later, silicon steel was massively produced in industry. In addition,
silicon improves the resistance to corrosion and elastic limit of iron. Altogether,
Fe-Si is probably the best alloy for low-frequency AC applications from magnetic,
mechanical, and economic points of view since it is composed only of cheap and
abundant elements.
The silicon content can be varied in a very large range and forms bcc solid
solutions up to 25 at.% Si, limit above which the alloys decompose into Fe3 Si
soft phase and a Fe7 Si3 hexagonal hard magnetic phase. Figure 8 shows the main
properties of Fe-Si alloys. By increasing the silicon mass content up to 10%, the
resistivity is multiplied by a factor 10, and the magnetization is reduced by 30%.
However, the increase of the hardness makes the material very brittle above 10
at.% ( 5 mass%). This increase in hardness also makes the rolling more difficult
and reduces lifetime of cutting tools. So, in practice (except the case of 6.5 mass%),
29 Soft Magnetic Materials and Applications 1453

Fig. 8 Silicon dependence of room temperature saturation magnetization and resistivity of Fe-Si
alloys. Coercivity is also given as a trend only since it depends on the manufacturing process

Si content is usually between 2 and 4 mass%. For low-end applications (household


and cars), low Si alloys are preferred to reduce the rolling and cutting costs. In
contrast, for permanently plugged systems (network transformers, big generators),
higher Si alloys with lower losses are preferred. Notice that most of Fe-Si alloys
contain aluminum to help lamination. Usually, Al is less than 1 mass% and has
effects on physical properties very similar to Si, except for conductivity which is a
little bit higher compared to equivalent amount of Si.

Processing
The mother alloys are produced from recycled silicon steels for a great part and
pure iron (produced from iron ore) and silicon. The alloy is decarburized by blowing
hydrogen in the melt and cast into a 20 cm by several meters slab. Then, it is reduced
down to few mm by several hot-rolling passes, followed by several cold-rolling
passes depending on the silicon content and the final thickness. At this stage, the
sheets are about 1 m wide in the form of a 1 m in diameter coil. The semi-processed
materials are annealed at 800◦ C for a few minutes and rolled again with a small
thickness reduction (skin pass). It is then sent to manufacturer for cutting and must
be annealed in three steps: 750°C for decarburization under partial H2 pressure,
800°C for grain growth, and then in a wet atmosphere for surface insulation by
oxidation. The fully processed materials are annealed immediately after cold rolling
at 800◦ C under partial H2 pressure for decarburization and then up to 1000°C for
1454 F. Mazaleyrat

grain growth and insulated by organic binder. The manufacturer can cut the sheets
and mount them without further treatment.

Texture
Cold rolling produces heavy deformation mainly by slipping in the most dense
plane. In a bcc lattice, the most dense plane is {110}. Accordingly, successive cold-
rolling pass tends to produce a {110} texture (grains have a {110} plane parallel
with the sheet plane). However, cold rolling is not perfectly anisotropic because all
passes are made in the same direction, called rolling direction (RD). Consequently,
a slight anisotropy in the plane is observed corresponding to a small {110}001
texture and/or fiber -type components (mainly {111}112 see, e.g., [35]). This kind
of sheets is called non-oriented silicon steels (NOSS).
Because 001 is the easy magnetization axis, permeability is stronger along RD
compared to transverse direction (TD). Starting from this fact, metallurgists tried to
improve the 001{110} texture in order to maximize the RD permeability. These
materials are called grain-oriented silicon steels (GOSS).

Designation of Magnetic Steels


There is no complete standard process for production of NOSS; if the general
scheme described above is followed, each manufacturer has its own process and
secrets. The result is that for the same grade, sheets from one provider or another
could have slightly different texture and properties.
The grade of magnetic silicon steels is defined by standards. The European
designation for magnetic alloys follows the general one for steels that starts with
a letter, here M for magnetic, followed by:

1. A number being hundred times total specific losses in watts per kilograms at a
frequency of 50 Hz and an induction of:
– 1,5 T for fully and semi-processed NOSS and regular GOSS
– 1,7 T for low losses and high induction (Hi-B) GOSS
2. A number corresponding to hundred times the thickness expressed in millimeters
3. A letter corresponding to the type of silicon steel:

A for fully processed NOSS


D for non-alloyed semi-processed steels
E for alloyed semi-processed steels
N regular GOSS
S low-loss GOSS
P Hi-B GOSS

For NOSS, standard thicknesses are 0.35, 0.5, 0.65, and 1 mm, but thinner gauge
sheets (down to 100 μm) are more and more proposed by steel companies. GOSS
are usually 0.35, 0.3, 0.27, and 0.23 mm thick, the last two being used generally for
Hi-B.
29 Soft Magnetic Materials and Applications 1455

Designation examples:
M 85-23 P Fe-Si grain-oriented sheet, 0.85 W/kg at 50 Hz under a 1.7 T sine
waveform peak induction, 0.23 mm thick.
M 240–35 A Fe-Si fully processed sheet, 2.4 W/kg at 50 Hz under a 1.5 T sine
waveform peak induction, 0.35 mmm thick
35 A 240 same in Japanese standard with permutation of numbers
36 F 145 same in the US standard, with a small difference in thickness (due to
older inch gauge) and 100×losses in W/lb at 60 Hz

Non-oriented Silicon Steel Sheets

NOSS are used in many different low-frequency applications, because of their good
mechanical and magnetic properties, in-plane isotropy, and low price. They are
proposed in a wide range of quality and price depending on the type of application
and the requirements in terms of performances.
In general, thicker gauges are cheaper and have higher losses and eventually
better permeability. This is illustrated in Fig. 9 where it is seen that thick-gauge
sheets have higher magnetization at low H and lower magnetization at high H (the
cross point is here about 300–400 A/m), but the 50 Hz losses are always higher for
thick gauges. As a consequence, they are used generally for low-power household
applications where price matters more than performance. In industrial applications,
the power is usually higher, and the cost of losses over the whole life of the device
has to be taken into account in the economic balance. As the size of the device
increases, heat is more difficult to release, so higher grade has to be used. If the
working frequency is higher than the network frequency, thinner gauge can be used
to reduce eddy current losses that becomes predominant above several hundred Hz.

Fig. 9 Comparison of different common grades of oriented and non-oriented Fe-Si sheets
1456 F. Mazaleyrat

Fig. 10 Power loss as a function of field angle with respect to rolling direction at 50 Hz (left).
Power loss at 1.5 T, 50 Hz as a function of thickness. Points are for higher grade material; line is
for eddy currents (right)

High-grade NOSS are used essentially in industrial electrical machines where the
flux is not everywhere parallel to lamination directions in stator teeth and rotating
in stator core.
The NOSS are considered isotropic in plane, but this is not exactly observed in
practice. At low field, losses are almost independent of the direction of the field with
respect to rolling direction (RD), though a little bit smaller at 0° as illustrated in the
polar plot (Fig. 10). At 1.4 T losses are quite similar at 0 and 90° , but they show
a marked maximum around 45°. For this reason, according to the standard, losses
have to be measured using strips cut in four different directions (0, 30, 60, and 90°).

Cutting
The fully processed sheets are delivered in the form of annealed and insulated sheets.
The user cuts it in the required shape using punching, water jet, or laser, but losses
are guaranteed only for Epstein format strips (300 mm × 30 mm). Because cutting
always affects locally the magnetic properties, users have to keep in mind that losses
could be higher in narrower zones such as stator teeth.
The semi-processed sheets have to be annealed after cutting. In this case, if the
user anneals cut parts in the conditions given by the provider, losses are guaranteed
in all parts whatever could be the cutting process or the size of cutting. However,
this annealing has a cost for the user, so because semi-processed sheets are cheaper
than fully processed, many electrical machine manufacturers use semi-processed
sheets without annealing. In this case, properties are no longer guaranteed and can
be significantly different from those claimed by the provider. Typical grades of semi-
process silicon steel range between M-50–340-D and M-65-1000-D.

Thin-Gauge Steels
In embedded power applications, in particular in electric vehicles, power density
of electrical machines is the main issue after energy storage. For this purpose,
29 Soft Magnetic Materials and Applications 1457

Table 1 Magnetic properties of Fe-Si alloys fully processed sheets [27]


Reference Si content ρ B(T ) HC 50
p1.5 400
p1.5 p12.5k
(mass%) (μcm) @ 2.5 kA/m (A/m) (W/kg) (W/kg) (W/kg)
M350-10-A 3.5 52 1.53 3.20 29.7 119
M270-20-A 3.5 52 1.59 2.4 29.7 205
M210-27-A 4.3 62 1.49 30 2.04 31.4 235
M235-35-A 4.1 59 1.53 35 2.25 41.2 249
M270-35-A 3.5 52 1.54 40 2.47 41.8 353
M330-35-A 2.7 42 1.56 40 2.94 57.9 517
M250-50-A 4.1 59 1.55 59 2.38 60.7 617
M400-50-A 2.7 42 1.59 50 3.57 91.7 899
M600-50-A 1.7 30 1.63 85 5.17
M800-50-A 1.0 23 1.65 23 6.60
M310-65-A 4.1 59 1.56 35 2.90
M800-65-A 1.3 25 1.65 100 6.74
M600-100-A 3.5 52 1.59 40 5.11
M1000-100-A 1.7 30 1.65 85 8.89

manufacturers increase the working frequency with a consequence on iron losses.


This led metallurgists to develop low-thickness NOSS. Figure 10 shows that
reducing the thickness obviously lowers the eddy current losses but increases the
hysteresis loss. For the standard point, the lower losses are obtained for a thickness
of about 0.25 mm. As the frequency increases above 400 Hz, advantage is clearly
for the thinnest gauge.
An overview of NOSS properties is given in Table 1.

Grain-Oriented Silicon Steel Sheets

By opposition to electrical machines for which the magnetic core should be


isotropic, the ideal material for transformers should be anisotropic with easy axis
corresponding to rolling direction (RD). Suitable texture should be either cubic
{100}001 or {110}001. The second one is known as Goss texture from the name
of its inventor, but coincidently, GOSS also means grain-oriented silicon steel. For
bcc crystals, the slip planes are highest density planes, {110}, so Goss texture is in
principle easier to obtain.
The advantage of using GOSS in applications where the flux is mainly directed
along the easy axis is clearly illustrated in Fig. 9; the induction at 800 A/m is almost
20% higher compared to NOSS and the loss at 1.5 T/50 Hz lower by a factor 4 to 10.

Conventional GOSS
Typical GOSS process starts by melting of an alloy containing about 3.2% silicon
and 0.03% carbon, followed by hot rolling down to 2 mm after heating to about
1458 F. Mazaleyrat

1300–1400 °C. Then the sheet is annealed and cold rolled with a reduction rate
of about 70% and annealed at 900 °C and cold rolled again with a reduction rate
of 50%. The coil is then decarburized in wet hydrogen and nitrogen at 800 °C,
coated by MgO, wound into a coil, and finally annealed between 1100 and 1200 °C
in dry hydrogen. The MgO coating is necessary to avoid sticking of the turns
during annealing, and it reacts with Si to form an Fe-Mg silicate at the surface,
a glass film also useful to ensure electrical insulation. The last thermal treatment
produces the secondary recrystallization yielding the texture if a proper growth
inhibitor (usually MnS) is used to impede primary recrystallization that would lead
to isotropic texture. The conventional GOSS have much improved permeability and
losses compared to NOSS if the field is applied in the RD, but it is degraded when
the field is applied perpendicularly. As a consequence, transformer cores should be
preferably built from straight parts rather than from EI cuttings (see Fig. 12 in such
a way the flux is always parallel to RD).

High Induction GOSS (HiB)


HiB sheets are manufactured using a process very similar to the conventional GOSS
except that cold rolling is realized in one pass only with almost 85% reduction and
that AlN is added to MnS as growth inhibitor. The result is an anomalous grain
growth of {110}001 crystals that can become very big (several cm). The magnetic
domains are also very big which produces large excess loss according to Pry and
Bean [29]. A common method to reduce dynamic losses is laser scratching. It
consists in making laser spot lines transversely to the RD (more or less one spot
each 0.1 mm and one line each 5 mm) that act as domain pinning points reducing
their extension to about 1 by 5 mm. In addition, more sophisticated glass films have
been developed with low dilatation coefficient – at least twice smaller than that
of Fe-Si (12 · 10−6 K−1 ). During annealing, internal stresses are released, but after
cooling the sheets remain under the tension of the glass film. Because of the positive
magnetostriction of Fe-Si, the domains tend to be thinner and better oriented toward
the RD, and lancet domains due to out-of-plane disorientation are suppressed (see
Fig. 11).

1 cm

Fig. 11 Effect of tensile stress along rolling direction (horizontal) on the domain structure . From
left to right, 4 MPa, 12 MPa, 24 MPa (courtesy of R. Barrué). Images are obtained by magneto-
optical Kerr effect microscopy
29 Soft Magnetic Materials and Applications 1459

Cutting
Conventional and HiB GOSS are mostly used in network frequency transformers.
So in most of cases they work at 50 or 60 Hz at high induction over 1.5 T for
conventional and over 1.7 T for HiB. The simplest way to build a transformer core is
to stack sheets cut in E and I shape (Fig. 12). The I is cut to make two windows in two
face-to-face E. EI cores are easy to realize: two stacks are made alternatively with E
and I (or several E/several I); the coils made separately are slipped in one stack, and
the other is slotted in. EI cores have two disadvantages: the flux is perpendicular to
RD in the yoke of the E, and for three-phase types, the window is bigger than the I
so there is always a waste. If low loss and low EMI is required, strip wound cores are
sometimes used in the ring or rectangular form. The single- or double-frame cores
are used for transformers above 1 kVA apparent power because the flux is mostly
parallel to RD except at the corners. To reduce further the losses, the joints are cut
at 45°and overlapped.
Different types of cuttings and grades are used depending on the power of the
transformer (Table 2):

overlaping
joint

E I core single frame core E I core double frame core


single phase stacked core 3 phase stacked core

Fig. 12 Different types of stacked cores for transformers and inductors

Table 2 Magnetic Standards ρ 50


p1.5 50
p1.7 B800 B2.5k
properties of grain-oriented
f (μcm) (W/kg) (W/kg) (T) (T)
Fe-Si alloys. pB is for the loss
M075-23-N 49 0.70 1.00 1.76 1.92
at induction B and frequency
f ; BH is the induction at field M083-27-N 46 0.80 1.17 1.76 1.92
H expressed in A/m [34] M090-30-N 46 0.87 1.24 1.74 1.91
M110-35-N 46 1.02 1.47 1.74 1.90
M090-20-S 49 0.61 0.88 1.80 1.93
M095-23-S 49 0.64 0.91 1.81 1.94
M110-27-S 46 0.74 1.03 1.81 1.94
M120-30-S 46 0.83 1.14 1.82 1.94
M135-35-S 46 0.91 1.26 1.83 1.96
M090-23-P 48 0.63 0.87 1.89 1.97
M095-27-P 48 0.69 0.93 1.90 1.98
M100-30-P 48 0.74 0.99 1.90 1.98
M115-35-P 48 0.87 1.14 1.90 1.98
1460 F. Mazaleyrat

– S < 1 kVA, EI core cut in conventional GOSS (low added value products)
– S < 1 kVA, strip wound core from thin conventional GOSS (high added value
products)
– 1 < S < 10 kVA, frame core cut in conventional GOSS
– 10 < S < 100 kVA, frame core cut in low-loss GOSS
– S > 100 kVA, frame core cut in HiB GOSS

GOSS are also used for inductors either using EI cores with air gap (in this case
layers are not alternated) or using cut strip wound rectangular cores. The latter,
often called C cores, are usually made from thin gauge (100 or 50 μm) and used
in power electronic devices up to 10 kHz. In power plant generators, significant
reduction of losses can be obtained using GOSS for the construction of the stator.
Parts containing several teeth are cut and put with RD radial, whereas yoke parts are
arranged with RD tangential.

Trends in Fe-Si Products

Improved Texture NOSS


The main drawback of {110} texture is related to the fact that this plane contains the
111 hard axes: when the magnetization turns from a 100 easy axis to another,
it should pass through a 110 direction, but these axes are out of plane so a strong
shape anisotropy is involved. The solution is to develop a {100} cubic texture or at
least to suppress a much as possible {111} components. For the same losses, the
induction at 5 kA · m−3 is about 1.7 T, viz., 0.1 more than the best fully processed
materials.

Low Si Alloys
Reduction of silicon content naturally leads to an improvement of the magnetization.
For example, reduction to 2% Si allows to reach 1.75 T at 5 kA · m−3 at the cost
of higher losses (above 3 W/kg at standard point). These products are laminated
with thickness between 0.5 and 1 mm for mechanical strength reasons. This larger
thickness combined with lower resistivity results in an increase in eddy current loss,
so they are limited to network frequency use (Table 1). Due to their larger ductility,
cutting cost is reduced, but because of larger magnetostriction, the properties are
more dependent on the cutting technique and annealing conditions.

Silicon-Enriched Electrical Steels


Looking at the magnetic properties of Fe-Si alloys, it is seen that if saturation
magnetization monotonically decreases, magnetocrystalline anisotropy and magne-
tostriction do so as well, and the latter is exactly zero for 6.5 w.% Si (12.14 at.%). As
a consequence, this alloy has a large permeability, low loss, and little sensitivity to
stress. If its saturation magnetization is lower than regular silicon steel, the induction
under a field of 2.5 kA/m is comparable with GOSS. However, this alloy is very
29 Soft Magnetic Materials and Applications 1461

Table 3 Magnetic properties of 6.5% Si steels [24]. The resistivity is not given for graded
materials since it is not constant
Type Thickness (mm) ρ(μΩcm) JS @ B(T ) HC p150 p1400 (W/kg) p11k
2.5 kA/m (A/m) (W/kg) (W/kg)
Graded 0.10 - 1.88 1.44 40 1.10 10.1 30.0
Graded 0.20 - 1.94 1.47 35 1.10 14.5 51.6
Uniform 0.05 82 1.80 1.38 26 0.70 6.50 18.8
Uniform 0.10 82 1.80 1.40 19 0.50 5.70 18.7

brittle and can’t be laminated. A solution was found in the early 1990s by the
Japanese company NKK (now JFE) introducing Si by chemical vapor deposition
(CVD). Starting from a low Si laminated sheet, annealing in an atmosphere
containing silane gas results in diffusion of Si into the metal. Using this method,
the obtained material exhibits excellent mechanical and magnetic properties. 6.5
silicon steel sheets can be made with two different grades: uniform and graded Si
diffusion. Because CVD is a diffusion process from the surface, the lamination is
enriched at the surface only, and then the silicon gradient can be controlled by the
annealing time. Longer annealing leads to homogeneous diffusion in the bulk. The
graded laminations have lower eddy current losses at high frequency because the
eddy currents flow mainly at the surface. In contrast, they have higher hysteresis
loss because there is a gradient of permeability and saturation magnetization across
the thickness (see Table 3.) Thus, the choice of homogeneous or graded material
depends essentially on the working frequency.

Iron-Cobalt Alloys

As shown by Pierre Weiss in 1912, alloying cobalt to iron has a spectacular effect
on magnetic properties since Fe-Co has larger magnetization than iron. This is
due to a modification of iron band structure, where the Fermi level is increased
in such a way that 3d-up sub-band is filled by electrons from 3d-down sub-band
(see  Chap. 15, “Metallic Magnetic Materials”).

Equiatomic Fe-Co Alloy

Fe2 Co has the strongest room temperature magnetization known but is relatively
hard. However, the equiatomic alloy, known as Permindur, has a small anisotropy
constant, a very soft behavior, and still a very high room temperature magnetization,
the reasons why it is preferred in spite of larger cobalt content and higher cost
(Table 4). In some applications, performance is more important than price of
the elements. In airborne systems, the weight of an equipment is paid by fuel
consumption at each takeoff and power of the turbine needed to provide potential
energy at 10,000 m from the ground. Under these conditions, a gain of 20% in
1462 F. Mazaleyrat

Table 4 Main characteristics of Fe-Co alloys. Magnetostriction is given for isotropic alloys, and
ultimate strength σU is given for the hard state (for annealed state, divide by a factor 2 to 3)
Composition Form t mm Density TC (°C) ρμcm JS T λS 10−6 σU MPa
kgm−3
Co49 Fe49 V2 Strips, bars 0.1–1.5 8120 900 40 2.4 70 1200
Co27 Fe73 Cr0.3 Mn0.5 Strips, bars 0.1–3.5 8000 980 20 2.4 40 1000
Co18 Fe79 Cr3 Strips, bars 0.1–3.5 7900 940 30 2.3 30 850

Table 5 Magnetic properties of Fe-Co alloys of 0.35 -mm-thick strips


Composition Annealing °C B(800 A/m) B(8 kA/m) HC 50
p1.5 p250 400
p1.5 p2400
T T A/m W/kg W/kg W/kg W/kg
Co49 Fe49 V2 750 2.05 2.28 120 3 5 44 84
850 2.1 2.28 70 2.3 3.8 42 74
Co25 Fe73 Cr2 750 1.47 2.02 200 6 9.5 70 140
850 1.47 2.02 100
Co18 Fe79 Cr3 700 >1.7 <2 150 5.5 6.7 77 99
900 >1.7 <2 260 3.3 4.3 66 88

induction results in a weight diminution of 20 to 30% and becomes economically


favorable. FeCo has four big drawbacks: a low resistivity (below 10 m), a
brittleness that make rolling very difficult, a bcc to fcc phase transformation about
1000 °C, and an ordering near 730 °C detrimental to magnetic properties and
making the alloy even more brittle. Among all possible substitutions, 2% vanadium
notably improves the resistivity, the ductility, and the stability, but high-temperature
purification annealing is still impossible.
Fe49 Co49 V2 is hot rolled down to about 2 mm, then heated and quenched in
water to avoid ordering, and finally cold rolled to a gauge between 0.35 and 0.1 mm.
In this form, it is essentially used for transformers and the rotor/stator of airborne
electrical machines, but it is also used in the bulk form for magnet pole pieces and
electromagnetic valves for car thermal engines. The DC coercive force is relatively
high compared to GOSS (see Table 5), but induction as high as 2.2 T can be reached
in a field of less than 10 kA/m and works at 400 Hz with an induction of 2 T.
Due to the large magnetostriction, it must be annealed to release stresses at low
temperature. Typically annealing just below the ordering temperature (750°) leads to
anisotropic magnetic properties adapted for electrical rotating machines. Annealing
at higher temperature (850°) reduces appreciably the losses but leads to ordering
in a texture detrimental to machines but favorable for use as transformer cores. Its
large magnetostriction (70 ppm) allows to use it as transducer or sensor in the lower
acoustic band.
In addition to the high magnetization, Fe-Co is also interesting for its high
mechanical strength. Indeed, in the near future, airborne generators are supposed
to turn at 15–30,000 rpm for a power of 150–200 kW, so centrifugal forces acting
on the rotor sheets will be critically high.
29 Soft Magnetic Materials and Applications 1463

Low-Cobalt Alloys

Because of the price of cobalt, users always try to reduce the cobalt content. Typical
low-cobalt alloys contain about 25 or 18%, to ensure saturation far above iron one
(i.e., 2.2 T). This is however paid by an increase in coercivity and losses as shown in
Table 5. These alloys have lesser performance compared to equiatomic ones but can
be used advantageously in several cases, for example, in landing gear motors which
work only several seconds per flight (so losses have no relevance) or in airborne
transformers because of the smaller magnetostriction producing less noise. Today’s
tendency is to improve the texture of low-Co alloys (easy axis parallel RD) for
transformer applications in order to reduce loss and effective magnetostriction.

Iron-Nickel Alloys

Iron-nickel alloys were studied for the first time by Hopkinson in the 1890s and by
others later for fundamental studies – and in particular by Guillaume who discovered
Invar – but it was only by the 1910s that Elmen from Bell Labs obtained Fe-Ni alloys
with permeability greater than that of pure iron. Invar is not properly speaking a soft
magnetic material as it is used only as a zero thermal expansion coefficient (CTE)
material. It is quoted here for historical reasons and because this feature is due
to compensation of thermal expansion and volume magnetostriction around room
temperature (above TC Invar has thermal expansion comparable to other metals).
Besides Invar, Fe-Ni alloys are of two kinds: bcc Fe-Ni (less than 30% Ni) and fcc
Fe-Ni (more than 30%). In general, there are no applications for the bcc form since
they do not exhibit better properties than Fe-Si at a higher price. By contrast, the
fcc form has many applications because a great variation in JS , μR , HC , TC is
observed with the Ni content. The most striking figure of these alloys is the strong
dependence of magnetocrystalline anisotropy on order. Indeed, when slowly cooled
after annealing in the range of 900–1200 °C, the alloy is ordered and K1 = 0
for 56% of Ni. In contrast, quenched alloys have K1 = 0 for 73% of Ni. As
a consequence, Ni-rich alloys are usually quenched and the Fe-rich ones slowly
cooled.
More detail on the fundamental aspects of Fe-Ni alloy magnetism can be found
in  Chap. 15, “Metallic Magnetic Materials”. An overview of industrial Fe-Ni soft
magnetic alloys is given in Table 6; they can be classified into three main families
described below.

Ni-Rich Alloys

These alloys are currently called permalloy after Elmen from Bell Labs in
1920, and their composition is always close to 80% Ni; for this composition,
the material undergoes a spin reorientation transition (say, a change of easy axis
from [111] to [110]) around room temperature and zero magnetostriction together
1464 F. Mazaleyrat

Table 6 Composition and structural properties of Fe-Co and Fe-Ni alloys


Reference Composition Form t (mm) Density TC ρ CTE
(°C) μcm (ppm/K)
Invar Ni36 Si0.25 Mn0.25 Fe63.5 Strip, 0.5–10 8130 230 80 1
bars
Mumetal Ni77 Fe14 Cu5 Mo4 Strip, 0.1–4 8700 400 60 12
sheet
Mo5-Permalloy Ni80 Fe15 Mo5 Strip, 0.1–1 8700 420 60 12
sheet
Mo6-Permalloy Ni81 Fe13 Mo6 Strip, 0.1–1 8700 420 60 12
sheet
36Fe-Ni Ni36 Fe64 Strip, 0.1–0.35 8100 230 75 1
sheet
40Fe-Ni Ni41 Fe59 Strip, 0.1–0.35 8150 330 60 4.5
sheet
50Fe-Ni Ni48 Fe52 Strip, 0.08–0.35 8200 450 60 8
sheet
Thermalloy Ni50 Fe41 Cr9 Strip, >0.4 8200 260 100 10
disk
Thermalloy Ni50 Fe40 Cr10 Strip, >0.4 8200 230 100 10
disk
Thermalloy Ni30 Fe65 Cr2 Cu3 Strip, >0.4 8150 120 88 6.6
disk
Thermalloy Ni30 Fe70 Strip, >0.4 8100 55 80 1
disk

in the disordered state. For this reason, the material properties are very sensitive
to the microstructure and order, so the heat treatment can change drastically the
magnetic characteristics such as coercivity and permeability. Typical heat treatment
consists in heating over 1000 °C, then cooling to 600 °C, and quenching in air.
This procedure retains disorder and promotes high permeability for Ni75 Fe25 .
Alternatively long-time annealing followed by slow cooling (1 day) produces
recrystallization and forms a Ni3 Fe-like ordered alloy. In this case, the maximum
permeability is obtained for lower Ni content. The drawback of Fe-Ni alloys is
the relatively low resistivity because the two elements are from the iron group and
have similar electronegativity. In industrial permalloys, 4 to 6% molybdenum is
usually added to increase the resistivity and to ensure minimum magnetocrystalline
anisotropy and magnetostriction. For example, 5% Mo substituted to Fe in Ni80 Fe20
results in an increase in resistivity from 20 to 60 ·10−8 m, but at the same time,
saturation magnetization is decreased from 1.05 to 0.8 T. Typical Mo-permalloy
contains 5% Mo and 80% Ni, and the magnetic properties can be changed by
the heat treatment: quenched, slowly cooled, recrystallized, or annealed in field.
Annealing is conducted in pure dry hydrogen usually between 1000 °C and 1200 °C
to remove impurities (C, O, S, etc.). The permeability can be enhanced by increasing
treatment temperature (less impurities, bigger grains). Rectangular hysteresis loop
is obtained by slow cooling promoting ordering. For this composition, the easy axis
29 Soft Magnetic Materials and Applications 1465

becomes [111] which (due to the sixfold symmetry) leads, in theory, to a remanence
ratio of 87%, very close to the experimental value. By opposition, flat loops are
obtained by annealing under a magnetic field applied perpendicular to the strip. A
first annealing is conducted at low temperature (about 500°) to adjust K1 = 0 and
followed by field annealing below TC . Mo5 Py is used in any application where
very high permeability and very low coercivity are necessary if magnetization is
not a dimensioning parameter (typically ground fault detector, current transformer).
Rings (toroids) realized by strip winding are the most used shape for permalloy, but
strip wound rectangular cut cores are often used for inductors. In applications where
both K1 = 0 and λs = 0 are required (typically for flexible magnetic shields),
the composition is modified to 6% Mo and 81% Ni. If higher Curie point and
magnetization are required, 4% Mo 80% Ni alloy is preferred (Table 6).
Mumetal is a variant of permalloy in which several % of copper are added
to improve the ductility. This reduces the cost of both lamination and forming
processes, so it is used generally for large rigid magnetic shields or zero field
chambers.

Fe-Rich Alloys

Iron-rich alloys contain at least 30% of nickel; they always have the fcc structure.
Equiatomic alloy is the most used because it offers a nice combination of high
magnetization and permeability together with low coercivity and losses. However,
this alloy exhibits relatively low resistivity which is detrimental to losses in high-
frequency applications (see Table 7). In this case, 36Fe-Ni may be used, but TC is
decreased by 100 K as it is closer to the Invar composition. Iron-rich alloys can be
processed by three routes:

Table 7 Magnetic properties of Fe-Ni alloys. Py is the usual abbreviation for permalloy; R and
F are for rectangular and flat hysteresis loops, respectively
Reference JS (T) HC (A/m) μi μmax BR /JS Loss (W/kg) λS (ppm)
f
static static static % pB
Mumetal 0.78 1.6 40 110 50 = 0.04
p0.5
Mo5-Py 0.80 0.8 75 200 50 50 = 0.02
p0.5 1
Mo5-Py R 0.80 0.8 75 200 85 1
Mo5-Py F 0.80 0.7 100 130 20 1
Mo6-Py 0.66 0.4 300 130 50 0
36Fe-Ni R 1.30 15 3 20 p150 = 1.1 20
36Fe-Ni 1.30 55 2 7
40Fe-Ni 1.45 6 66
50Fe-Ni 1.60 8 4 35 50 p150 = 0.1 24
50Fe-Ni R 1.60 9 100 95
50Fe-Ni F 1.60 1.5 40 100 30
1466 F. Mazaleyrat

1. Moderate hardening (60% skin pass) and 1100 °C annealing under H2 lead to
isotropic texture. This treatment results in an isotropic magnetic behavior and
low losses. The permeability is moderate compared to Ni-rich alloys but much
higher than NOSS. An induction of 1 T is obtained in a field of only 10 A/m, an
order of magnitude lower than NOSS or FeCo. These features are ideal for small
and efficient electrical machines, when the power density is not a key issue. They
are currently used in the form of stamped sheets for stators or rotors and in the
form of bulk parts in electromagnetic relays.
2. Strong hardening (>90%) and >1100 °C annealing produce an anisotropic
cubic texture {100}001, leading to very high rectangularity. Growth inhibitors
(selected impurities) have to be added in the alloy to stabilize the texture and
avoid secondary recrystallization which slightly increases the coercivity. They
are suitable for low-loss transformers, especially at 400 Hz or above, as the
permeability is very high, and in applications where high rectangularity is
needed such as magnetic amplifiers or different types of sensors (fluxgate; see
 Chap. 31, “Magnetic Sensors”).
3. Secondary recrystallization at high temperature (>1100 °C) of high-purity alloy
results in anomalous growth of crystals. Subsequent field annealing just below
TC produces uniaxial anisotropy in the direction of the applied field. 55Ni alloy
is preferred because K1 is appreciably smaller in the ordered state compared to
alloy containing 48% of Ni. The field-induced anisotropy is of order 400 Jm−3
which is a bit stronger than the magnetocrystalline one of the ordered alloy (about
200 Jm−3 ) and sufficient to flatten the loop. The remanence to saturation is low,
so the material demagnetizes itself: this feature associated to high induction is
interesting for unipolar transformers (the maximum induction swing is ΔB 
JS −BR ) or in any application similar to that of their Ni-rich counterparts (current
transformers, energy meters, etc.) but with smaller price due to the smaller Ni
content.

Thermal Alloys

Thermal alloys have been originally developed to compensate the effect of thermal
fluctuation in magnetic circuits. These alloys have usually a composition close to
Fe30 Ni70 which is at the limit between bcc and fcc phase and has a Curie point of
50 °C. Cr is generally added in the composition since it moves the bcc/fcc limit to
lower Ni content and offers an additional parameter to control TC together with a
reduction the price of the elements.
This is particularly useful for permanent magnet circuits which are usually quite
sensitive to temperature. When a highly stable induction in the gap of a magnetic
circuit containing a permanent magnet is required, a thermal alloy part is used as a
shunt to the magnet. It was used in the past in electrical measurement instruments,
tachometers, relays, etc., but today most of these applications are obsolete. Recently,
a new class of applications appeared for controlled Curie point materials, i.e.,
induction cookware. In this application, a magnetic layer is needed to concentrate
29 Soft Magnetic Materials and Applications 1467

the flux in the pan bottom and maximize the energy transfer. Associated to the
copper layer (needed to ensure thermal homogeneity and strong eddy currents) and
the stainless steel pan body, the alloys must have a similar dilatation coefficient to
the other materials. Advantageously, the Curie point may be adapted to naturally
regulate the temperature to avoid burning the food but also for safety as TC can be
tuned below the ignition point of oil. For frying pans, the Curie point is 260 °C,
whereas for other pans it can be lower (230 °C) or only just above boiling point of
water (120 °C). Eventually, this natural regulation reduces the electrical stress on
the inverter, limiting regulation switching.

Amorphous and Nanocrystalline Alloys

The amorphous state of matter, or glassy state, is a solid state having no long-
range order. In a first approach, amorphous solids are considered to have tetrahedral
local order. An amorphous state of matter is natural for silicates formed by rapid
cooling of lava to form obsidian or volcanic glass. However, only metallic glasses
are magnetic, and pure metals (or transition-metal alloys) cannot be obtained in
the glassy state. Thus, some elements called glass formers must be added, and the
quenching rate must be of the order of magnitude of 106 K/s.1 The glass former must
have smaller atomic radius than transition metals, and, eventually, an addition of a
bigger atom (refractory metal) can help. To be magnetic, a glass must contain more
than 60% of transition metals carrying magnetic moment. The general formula is
T1−x−y Mx Ry where for industrial material:

• T = Fe, Ni, Co, a transition metal


• M = Si, B, C a metalloid with 20 < x < 30 at.%
• R = Nb, Mo a refractory metal with y < 4 at.%

At the industrial scale, the alloys are prepared from pure elements except boron
provided by F3 B or F2 B compounds. The elements are usually melted by induction
in vacuum and poured in a crucible maintained at a temperature several degrees
higher than the melting temperature (1100–1300°C). The liquid metal is projected
by inert gas into a nozzle fitted with a thin slit very close to the rim of a cold wheel
turning with a peripheral speed of about 30–40 m/s. This process is known as planar-
flow casting. The rim is usually made of copper – or copper alloy, e.g., Be brass – to
ensure efficient thermal transfer and water cooled. If the parameters, speed, pressure,
and temperature are properly tuned, the alloy is directly transformed into a ribbon
having a width equal to that of the slit and a thickness of about 20 μm. Depending
on the application, the thickness can be tuned between 18 and 40 μm and the width
from several mm to over 100 mm.
The as-quenched ribbons are in the amorphous state, but they have to be annealed
to release quenching stresses and optimize the magnetic properties. In the case of
cobalt-based alloys, annealing also partly releases free volume. Because iron-based
alloys do not retain free volume, annealing results in a significant embrittlement.
1468 F. Mazaleyrat

Table 8 Composition and structural properties of metallic glasses and nanocrystalline ribbons
[22, 23, 37]
Reference Composition t (μm) TX Density E α
(°C) (MPa) (ppm/K)
Metglas 2605A1 Fe78 B13 Si9 25 550 7.18 <110 2.0
Metglas 2605S3 Fe77 Cr2 B16 Si5 20 535 7.29 <110 6.7
Metglas 2605SC Fe81 B13.5 Si3.5 C2 20 480 7.32 <110 5.9
Metglas 2605CO Fe66 Co18 B15 Si1 25 430 7.56 <110 8.6
Metglas 2826MB Fe40 Ni38 Mo4 B18 20 410 7.90 <110
Metglas 2714A Co66 Fe4 Ni1 B14 Si15 18 550 7.59 <110
Metglas 2705M Co69 Fe4 Ni1 Si12 Mo2 B12 20 520 7.80 <110
Vitrovac 6025 Co66 Fe4 B12 Si16 Mo2 20 530 7.70 <150
Vitrovac 6155 CoBSi 20 480 7.84 <150
Vitrovac 6030 Co70 (Fe, Mo)2 Mn5 (B, Si)23 20 480 7.60 <150
Vitrovac 6006 Ni, Co, Fe)8 0(B, Si)20 20 430 7.80 <150
Vitroperm 800 Fe73.5 Si15.5 B7 Nb3 Cu1 20 630 7.35 190
Vitroperm 500 Fe74.5 Si13.5 B8 Nb3 Cu1 20 630 7.35 190
Finemet Fe73.5 Si13.5 B9 Nb3 Cu1 18 630 7.40 180
t is the thickness, TX the crystallization temperature, E Young’s modulus and α the dilatation
coefficient

Annealing is usually performed at about 350–400° for cobalt-based and 400–450°


for iron-based alloys, preferably 100 K below the cristallization temperature (see
Table 8).

Iron-Based Amorphous Alloys

The iron-based amorphous alloys exhibit the highest magnetization of all amor-
phous materials due to the large iron content, about 1.6 T (eventually, magnetization
can be improved up to 1.8 T by cobalt addition). As a consequence, they are par-
ticularly fitted for low and medium frequencies when high magnetization is the first
dimensioning parameter. Because of the long-range disorder, the magnetocrystalline
anisotropy is theoretically zero, but the magnetostriction is usually quite strong
(about 25 ppm). This strong magnetostriction has been sometimes used for sensors
or transducers, but in general it is detrimental to the magnetic properties because
internal stresses cannot be totally released, so the permeability remains within the
order of magnitude of 105 “only” and the coercive field is never below several A/m.
In the 1980s, a large variety of iron-based amorphous alloys have been developed,
in particular by Metglas:

• SA1, for low-frequency applications and in particular network transformers


• S3, for high-frequency power electronics
• SC, for magnetoelastic-based applications (transducers and sensors)
29 Soft Magnetic Materials and Applications 1469

Table 9 Magnetic properties of metallic glasses and nanocrystalline ribbons. M is for Metglas
( [22]), VP for Vitroperm (Vacuumschmelze [37]), FT for Finemet (Hitachi [23]), and NP for
NanoPhy (Aperam [5])
Reference JS (T) HC μmax BR /JS Loss TC (°C) ρ (Ωm) λS
(A/m) static % (W/kg) (ppm)
static 100 kHz,
0.2 T
M 2605A1-Z 1.56 2.4 40,000 90 500 415 1.3 27
M 2605A1-F 1.56 2.4 6000 5
M 2605S3-A 1.41 4.8 35,000 50 90 358 1.38 20
M 2605SC 1.61 3.2 30,000 600 370 1.25 30
M 2605CO 1.8 4 25,000 4 90 145 1.23 35
V 6006 0.42 500 95 150 120
M 2826MB-Z 0.88 0.4 80,000 90 500 353 1.38 12
M 2826MB-F 30,000 5
M 2714A 0.55 0.3 115,000 95 50 205 1.42 0
M 2714A 0.55 0.3 30,000 95 45 205 1.42 0
M 2705M 0.77 1.2 600 * 100 365 136 <1
V 6025-Z 0.55 0.3 1000 95 50 210 140 0.2
V 6025-F 100,000 5
V 6155-F 1 0.8 1000 5 58 485 110
V 6030-Z 0.82 1 300,000 95 80 365 130 0.2
V 6030-F 3500 5
VP 800-F 1.2 0.5 100,000 8 40 600 1.15 0.5
VP 500-F 1.25 0.5 20,000 4 40 600 1.15 0.5
FT-3H 1.23 0.6 30,000 89 80 580 1.35 2.3
FT-3M 1.23 2.5 70,000 50 40 580 1.35 2.3
FT-3L 1.23 0.6 17,000 5 34 580 1.35 2.3
FT-3M 1.23 2.5 70,000 40 40 580 1.35 2.3
NP Kμ 1.27 6 200-3000 1 100 580 1.35 <1

• CO, for same applications as SA1 but with higher induction. The magnetic
properties of these alloys are summarized in Table 9

However, today, only Metglas 2605SA1 is produced as iron-based alloy. They


are used essentially in network transformers. Compared to GOSS, they have much
smaller working induction but also much smaller losses (Fig. 13). Transformers
made from amorphous cores are more expensive because of a larger amount of
magnetic material required and more complicated construction, but as they save
unload losses, they are economically viable considering the cost of losses integrated
over 30 years life. Network transformers rating over 100 kVA are currently
developing owing to the increasing cost of energy. Most of them are five-limb type
made from four rectangular rolled open frames. Properties of Metglas SA1 have not
been improved significantly since the 1980s; only recently Metglas has proposed a
1470 F. Mazaleyrat

Fig. 13 Left: hysteresis loops of Fe-based Metglas SA1 samples annealed in longitudinal and
transverse field. Right: comparison of magnetization and iron loss curves of Metglas SA1 and HB

new alloy with improved induction (HB) obtained by using a larger amount of Fe
(about 85 at.%) but still not available.
Iron-based amorphous metals are used also in the form of C-cores cut from rolled
rectangular frame or powder cores.

Cobalt-Based Amorphous Alloys

Due to the smallest magnetic moment of cobalt atoms compared to iron, cobalt-
based alloys have markedly lower magnetization than iron-based ones (see Table 9).
In contrast, because they have a very low magnetostriction and low internal stresses
(as they retain free volume), they have the smallest coercivity of all magnetic
materials known and the highest permeability (see Table 9). In addition, thanks
to this nearly zero magnetocrystalline and magnetoelastic anisotropy, uniaxial
anisotropy can be induced by field annealing resulting in perfectly square or
perfectly linear hysteresis loop (Fig. 14 left box). These features are particularly
interesting for saturable cores or when both high permeability and linearity are
required. Because the cobalt is expensive, they are limited to high value-added
applications (sensors) or to applications where a small quantity of material is
required (anti-theft tag).

Nanocrystalline Alloys

Nanocrystalline alloys, known as Finemet (Hitachi trademark), have been discov-


ered in 1988 from iron-based amorphous alloys of specific composition [40].
The amorphous ribbon is produced by the same process as others, except that
the annealing temperature is pushed to 530–550°C to partially devitrify the alloy.
29 Soft Magnetic Materials and Applications 1471

Fig. 14 Hysteresis loops of longitudinal and transverse field-annealed cobalt-based amorphous


alloy (V6025). Left, iron loss as function of induction for various frequencies in Co-based Metglas
2714, Finemet, and Fe-based Metglas SA

Copper forms very small clusters playing the role of nucleating agent for Fe-Si
particles, whereas Nb acts as a growth inhibitor. The result is a homogeneous
distribution of 10–15 nm Fe80 Si20 particles representing about 65 vol.% in the
optimal state, embedded in the remaining amorphous matrix [33]. Owing to the
small grain size, the anisotropy is averaged over the exchange length resulting in a
nearly zero anisotropy following the mechanism proposed in [2] and demonstrated
conclusively in nanocrystalline materials by Herzer [20]. In addition, because
the Fe80 Si20 nanocrystals have negative magnetostriction, whereas the amorphous
remaining phase (composition close to Fe70 B30 ) has a positive one, proper tuning
of the Si content and the annealing leads to compensation down to nearly zero [21].
As a consequence, Finemet-type nanocrystalline materials have advantages of both
iron-based and cobalt-based amorphous, namely, high magnetization and extreme
softness (see Table 9). As cobalt-based metallic glasses, longitudinal field annealing
produces squares loops and transverse field annealing flat loops with high linearity
and low squareness (Fig. 15 left box). Compared to amorphous alloys, Finemet
exhibits loss much lower than Fe-based alloys and equal to Co-based at lower price
and possibly higher induction (Fig. 14 right box).
The most striking figure of this alloys is the effect of stress applied during
annealing. As shown firstly by Glaser [19], stress annealing induces a very strong
anisotropy perpendicular to the tension direction and directly proportional to the
tension applied. This results in very flat and linear loops; those permeabilities can
be directly controlled by the stress according to μR 7 × 104 /σ where the stress
σ is expressed in MPa [3]. Stress annealed strips are currently produced using
Arcelor-CNRS process [4] and usually rolled into ring shape. These components
have the capacity to store energy, not in an air gap as usual but in the material itself.
1472 F. Mazaleyrat

Fig. 15 Hysteresis loops of Finemet-type nanocrystalline alloys (labels indicates permeability).


Left transverse field-annealed material from Vacuumschmelze, right, stress annealed samples from
Aperam (Kμ). Note the two orders of magnitude on x scale

The advantage of stress annealed nanocrystalline cores compared to gapped core


or powder cores is a much higher linearity, lower losses for same energy stored,
and the absence of flux leakage. Among all soft magnetic materials, Finemet-type
nanocrystalline alloy is from far the most versatile since its permeability can be
tuned from 200 to 100 000 depending on the field or stress applied during the
nanocrystallization heat treatment (Fig. 15 right box). As a consequence, they can be
utilized in a wide range of applications where a high permeability is required with
a flat or square loop as sensors, filters, and transformers, where a low permeability
and high linearity is required as chokes, inductors, sensors, filters, etc.

Soft Ferrites

Spinel Ferrites

Spinel ferrites are iron oxides having the crystal structure of the mineral MgAl2 O4 .
The general formula of ferrites is MFe2 O4 where M is a divalent metal or a
combination having an ionic radius compatible with available octahedral (B) and
tetrahedral (A) sites of the lattice. Depending on the ionic diameter of the divalent
metal, the spinel can be normal (divalent metal in A site, like Zn) or inverse (divalent
metal in B site, like Fe, Co, Ni) or sometimes mixed (like Mn or combination of
two divalent ions or even a combination of univalent and trivalent ions). Actually
the position of M2+ depends both on its diameter (big ion goes in bigger B site)
and electrostatic interactions (more charged ion goes in the B site). Because all
magnetic exchange interactions are antiferromagnetic, the energy of the lattice is
usually minimized for the minimum magnetic moment. Consequently, because Fe3+
has the largest magnetic moment (5 μB ), Fe3+ moments cancel. In simple ferrites
29 Soft Magnetic Materials and Applications 1473

Table 10 Distribution of ions in spinel sites for simple ferrites and example of a mixed ferrite
Normal Inverse Zn0.25 Ni0.75 Fe2 O4
(A) [B] (A) [B] (A) [B]
2+
 3+  3+
 3+ 2+  3+

 3+ 
M Fe2 O4 Fe Fe M O4 Zn 3Fe 5Fe 3Ni2+ O16
tetra octa tetra octa tetra octa
↑ ↑↓ ↓ ↑ ↑ · ↓↓↓ ↑↑↑↑↑ ↑↑↑

(when M is one element), the net magnetic moment per formula is given by that of
M2+ , as illustrated Table 10. The situation is a little bit more complicated for mixed
ferrites but very interesting when M is a combination of Zn and another ion holding
a magnetic moment. Because Zn goes in (A) and because it has a zero magnetic
moment, the moment associated to (A) is reduced, and if Zn content is smaller than
0.5, the magnetic order keeps that of the inverse spinel so the net moment increases
(see Table 10). This is often used in commercial ferrites in the aim of enhancing the
magnetization. However, increasing the Zn content over 0.5 yields partial normal
spinel magnetic order and thus a reduction of the magnetization. The reader would
find more detail in  Chap. 17, “Magnetic Oxides and other Compounds”.
In spite of a strong magnetic moment, iron ferrite – aka magnetite, load stone –
is not used as a soft magnetic material because of its relatively high coercivity.
Fe2+ has a nonzero orbital moment that leads to strong spin-orbit coupling
and consequently strong spin-lattice coupling (in other words magnetocrystalline
anisotropy). Another drawback of magnetite is its conductivity. Indeed, the Fe2+
is the octahedral site, which is favorable from ionic diameter point of view but
unfavorable from electrostatic point of view, so hopping of an electron from Fe2+ to
A site Fe3+ yields little change in energy, and this hopping is made easier by thermal
excitation. As consequence, magnetite is an intrinsic semiconductor; its conductivity
is typically 10−4 m at room temperature. It is thus important for high-frequency
application to replace totally all ferrous ions. Technically speaking, hundred kinds
of ferrites can be (and have been) made since elements from Ti to Zn plus Mg, Li,
Al, etc., can be substituted. Fortunately, Darwin’s theory also applies to technology
so few families only of ferrite have markets, mainly Mn-Zn and Ni-Zn and, in a
much smaller extent, Li and Mg ferrites.

Synthesis of Sintered Ferrites

Raw Materials – industrial ferrites are produced by the ceramic route. In general,
the precursor materials are simple oxides: Fe2 O3 , MnO, ZnO, NiO, CuO, CoO, etc.
The purity and the grain size of the powders have significant influence on the final
properties. Ideally, they should be fine and homogeneous in grain size to ensure the
chemical homogeneity.
Milling – after weighting the oxides as a function of the expected composition,
the powder is milled in order to get a homogeneous mix with finer grains. This
1474 F. Mazaleyrat

operation is important to optimize the reaction. Usually the mill is made from
stainless steel so excess iron may be brought at this stage.
Firing – the mix is heated in air at a temperature between 900 and 1100°C,
depending on the grain size, to produce the solid-state reaction. At this stage, the
spinel phase is formed.
Milling – because firing produces grain growth, a second milling is necessary to
refine the grains, preferably below 1 μm, before sintering. At this stage, different
additives can be introduced in small quantities such as V2 05 , to reduce sintering
temperature, Si02 , to limit grain growth, CaO, to increase the resistivity, etc.
Forming – the powder is now mixed with an organic binder (usually PVA) and
pressed in a mold to give the final shape of the core. There is a great variety of
shapes that can be made industrially: rings, E, I, U, pots, rods, screws, plates, tubes,
etc. Pressing is generally uniaxial; isostatic pressing is used only at laboratory scale
to realize big cores or with complex shapes.
Sintering – To densify the core and to obtain good properties, sintering must
be conducted with caution. Depending on the composition, the grain size before
sintering, and the expected grain size after, the temperature is usually above 900°C
and up 1350°C. Sintering takes several hours, and cooling rate must be low,
particularly for bigger cores, in order to avoid cracking. NiZn ferrites are sintered
in air, but MnZn ones must be sintered in controlled atmosphere. Indeed, the partial
oxygen pressure in air is such that it would oxidize Mn2+ , so it must be controlled
according to ln p02 = −K/T .

Mn-Zn Ferrites

Manganese simple ferrite is a mixed type one: Mn ions are found in both sites.
Mn2+ substitution is very interesting because this ion exhibits a strong spin (5/2)
and a zero orbital moment. Thus, Mn-based ferrites can be extremely soft.

Influence of Chemical Composition on Intrinsic Magnetic Properties


The magnetic properties can be easily modulated by tuning the Mn-to-Zn ratio. Zn
substitution has three main effects on properties:

– Enhances the magnetic moment (within the limit of 0.5 per f.u.) as said before
– Affects AB exchange coupling resulting in a decrease in Curie temperature (TC )
– Reduces the magnetocrystalline anisotropy (K1 ) and thus tends to increase the
permeability (μR )

Interestingly, because higher magnetic moment can be obtained only at the price
of a reduction in Curie temperature, the highest magnetization at room temperature
is obtained for a relatively small Zn content (x 0.15).
Mn ferrite has a low magnetocrystalline anisotropy (due to 3d5 configuration of
ions) with a [111] easy axis, whereas Fe ferrite has strong anisotropy (due to Fe2+ )
with a [110] easy axis. As a consequence, a controlled excess of Fe can produce a
29 Soft Magnetic Materials and Applications 1475

Fig. 16 Permeability and loss dependence of high-frequency MnZn power ferrite on temperature

Table 11 Main applications of Mn1−x Znx Fe2 04 and typical compromise on Zn content
Application Main parameter Zn TC (°C) μ0 MS (T) μ0 MS (T)
content 0K 300 K
Telecom transformers MS 0.2 250 0.7 0.54
Power transformers TC , MS 0.3 230 0.78 0.5
Wide-band transformers μR , TC , loss 0.5 150 0.9 0.4
Signal transformers μR 0.55 130 1 0.39

spin reorientation transition (SRT) revealed by a maximum in permeability, and the


SRT temperature will depend on the Fe2+ rate (usually less than 0.1). Fine-tuning
of this rate can maximize permeability at working temperature or make it stable on
a wide range of temperature. Figure 16 is showing a typical thermal dependence of
the permeability of a MnZn ferrite. Here the Zn and Fe2+ rates are tuned to get a
SRT at 100 °C as revealed by the second Hopkinson peak.
Thus, choosing the composition of a MnZn ferrite is a compromise depending on
the most dimensioning parameter of the application. Typical examples are given in
Table 11.

Influence of Microstructure on Extrinsic Magnetic Properties


The permeability, coercivity (HC ), or losses are microstructure-dependent proper-
ties. When the permeability is related mostly to domain wall motion, it is usually
accepted as a first approach that

1 JS D
μR ∝ ∝√ (47)
HC K1
1476 F. Mazaleyrat

where D is the grain size. So getting the highest permeability means large Zn
content and coarse grains. However, MnZn ferrite, like magnetite, has a relatively
low resistivity (even without ferrous ions) due to electron hopping between divalent
and trivalent Mn ions. The intrinsic resistivity of MnZn ferrite is about 0.1 m
so in high-frequency applications eddy currents in the core would produce high
losses (see section “Eddy Currents”). CaO is added for this purpose because it’s an
electrical insulator and because the ionic radius of Ca2+ is too big to enter in the
spinel lattice. Thus, after sintering CaO segregates at grain boundaries forming an
insulating barrier of thickness t. The electrical resistivity can be written in a first
approximation ρeff (t + D) = ρCaO t + ρMnZn D. As ρCaO t  ρMnZn D and t and
D are several nm and μm, respectively:

t
ρeff ρCaO (48)
D

On the other hand, this insulating layer is paramagnetic, so the permeability is


reduced by the air gap:

μR
μeff = (49)
1 + μR Dt

As a consequence, in MnZn ferrites the increase in resistivity is always at the


cost of a drop in permeability. A side effect of the CaO barrier is grain growth
and densification inhibition resulting in larger porosity. As a first approximation, if
1
the material is porous, the factor Dt in (49) is changed to Dt + d − 3 − 1 where d is
the relative density. As a consequence, CaO alone reduces strongly the permeability.
This effect is usually counterbalanced by addition of a flux. A flux is an oxide having
a melting point lower than the sintering temperature: the liquid phase enhances
strongly the diffusion and in turn favors the grain growth and density. For this
purpose, SiO2 and V2 O5 are mostly used (by less than 1% of the mass), but
everything depends on the customs of the manufacturer.
In Table 12, properties of ferrites for broadband transformer typical application
are presented. They have approximately the same Mn/Zn ratio (about 1/2), and a
very good correlation is found between resistivity and permeability using (49) and
(48). Power ferrite have usually a smaller Mn/Zn ration (about 1/3) to increase TC
and a thicker insulating barrier to improve the resistivity. For these two reasons,
they have a lower permeability (see Table 13). A typical example of permeability
and loss dependence on temperature is given in Fig. 16. Because high-frequency
power applications require low loss, the insulation layer must be thick enough. To
compensate the decrease in permeability, the SRT is tuned to get a high permeability
at the working temperature (here 100 °C). Another consequence of the anisotropy
cancellation is the strong variation in losses with temperature. Thermal stability
is good on the low temperature side of the SRT since losses decrease with T ,
but, in contrast, on the high temperature side, losses increase with T which in
turn contribute to increased temperature. As a consequence, if the cooling is not
29 Soft Magnetic Materials and Applications 1477

Table 12 Main Mn1−x Znx Fe2 04 materials used for broadband transformers, pulse transformers,
common mode chokes, and delay lines. References starting with a letter are from TDK, with a
number from Ferroxcube
Ref. x B1k (T) B1k (T) μi fR (Hz) HC (A/m) TC (°C) ρ (m) fmax (Hz)
(approx) 25°C 100°C ± 30% 25°C
T46 0.55 0.40 0.24 15000 400 k 7 130 0.01 100 k
T38 0.55 0.43 0.26 10000 300 k 8 130 0.1 100 k
T36 0.55 0.40 0.24 7000 300 k 22 130 0.2 100 k
T37 0.55 0.38 0.24 6500 700 k 9 130 0.2 300 k
N30 0.55 0.38 0.24 4300 1M 12 130 0.5 400 k
3E6 0.55 0.39 0.22 12000 400 k 5 130 0.1 100 k
3E7 0.55 0.39 0.22 15000 250 k 4.5 130 0.1 100 k

Table 13 Main Mn1−x Znx Fe2 04 materials used for power transformers and power electronics.
References starting with a letter are from TDK, with a number from Ferroxcube.B1k is the
induction measured at 1 kAm−1
Ref. x B1k (T) B1k (T) μi fR (Hz) HC (A/m) TC (°C) ρ (m) fmax (Hz)
(approx) 25°C 100°C ±30% 25°C
N72 0.37 0.48 0.37 2500 20 M 23 210 12 300 k
N27 0.35 0.50 0.41 2000 15 M 33 220 3 150 k
N49 0.3 0.49 0.40 1500 30 M 42 240 17 1M
3C93 0.3 0.52 0.43 1800 20 M 13 240 5 300 k
3F35 0.3 0.5 0.42 1400 50 M 40 240 10 1M

sufficient, there is a risk of thermal runaway. Losses of selected ferrites are given in
Annex 1, Table 17, and compared with competing materials (Table 13).

Ni-Zn Ferrites
The simple nickel ferrite is an inverse spinel: this means that the Ni2+ ion goes in the
octahedral site only. The effect of Zn substitution on intrinsic magnetic properties is
very similar to that in MnZn ferrites, except that Ni gives lower magnetic moment
(2 μB ) and higher Curie point and anisotropy constant than Mn. This means that
NiZn ferrites can be used at higher temperature than MnZn ones in general and
that their permeability will be lower. However, the great advantage is that Ni can
be divalent only and iron trivalent only, so conduction by electron hopping is
impossible. As a consequence, the resistivity is within the order of magnitude of
>105 m. A very small excess of iron results in a spectacular drop in resistivity
(six orders of magnitude); it is thus usual to prepare the ferrite with a slight iron
default (e.g., 1.96 instead of 2) in order to compensate iron pollution from steel
balls and vial of the mill. NiZn ferrites can be sintered in air (no control of the
oxygen pressure needed) without prejudice of the magnetic or electrical properties.
As a consequence, there is no need to introduce interfacial resistive oxide. The
magnetic properties of NiZn ferrites are controlled by two parameters, the zinc
content and the microstructure. As in MnZn ferrite, the Zn lowers the Curie point
and by consequence the magnetocrystalline anisotropy and in turn enhances the
1478 F. Mazaleyrat

Table 14 Main Ni1−x Znx Fe2 04 materials. References starting with a letter are from TDK, with
a number from Ferroxcube. B5k is the induction measured at 5 kAm−1
Ref. x B5k (T) B5k (T) μi fR (Hz) HC (A/m) TC (°C) ρ (m) fmax (Hz)
(approx) 25°C 100°C ± 30% 25°C
K1 0.30 0.31 0.28 80 50 M 400 425 105 12 M
4C65 0.35 0.38 0.34 125 55 M 280 350 105 1M
4B1 0.38 0.36 0.31 250 25 M 150 260 105 1M
4A11 0.60 0.34 0.23 850 7M 33 180 105 1M
K10 0.65 0.32 0.24 800 4M 165 115 105 1M
4A20 0.72 0.26 0.15 2000 1.4 M 11 100 105 300 k

permeability. In Table 14, a clear correlation is seen between TC and μi . Also the
sintering temperature plays a role since it affects the grain size and the density (49).
For example, in order to have a large bandpass, 4C65 has low zinc content and is
sintered at lower temperature to get low density (4500 kg · m−3 , 86%) and small
grain size. In contrast, to get a large permeability, 4A20 has high content of Zn
and is sintered at higher temperature to get high density (5000 kg · m−3 , 96%) and
coarser grains.

Other Soft Ferrites

Ni-Zn-Cu Ferrite
This family of ferrite is basically issued from NiZn one. It was developed for the
manufacturing of micro-inductances. A micro-inductance is a component made by
co-sintering of a metallic coil embedded in ferrite (typically from 1 × 0.5 × 0.5 mm
to 2 × 1.25 × 1.25 mm). This is realized by stacking bonded ferrites and silica tapes
on which coils are printed. In principle, noble metals (Pd, Pt) are compatible with
ferrites, but they are too expensive. As a consequence, the sintering temperature is
decreased by substitution of copper to Ni to be compatible with silver (about 950°C).
The typical formula is Ni1−x−y Znx Cuy Fe2 04 , where 0.3 < Zn < 0.6 and 0.1 <
Cu < 0.2. In principle, manufacturers provide properties of the components but not
that of the material. However, properties of these ferrites are comparable to that of
NiZn, i.e., a permeability in the range of several hundreds with substantially lower
magnetization [25]. Also Ni-Zn-Cu ferrites show very low loss at high frequency
(about 300 mWcm−3 at 1.5 MHz, 25 mT) when a small amount of Co is substituted
to Ni in order to produce a SRT close to room temperature [31, 32].

Microwave Ferrites
The term microwave usually refers to UHF to mm waves (300 MHz to 300 GHz
according to IEEE standard), but ferrites are used only below 110 GHz (W-band).
In microwave applications, the main quantity characterizing the material is not the
permeability or the magnetization but the gyromagnetic resonance frequency. In soft
29 Soft Magnetic Materials and Applications 1479

magnetic materials, the anisotropy field is by definition low, so the main field is the
dipolar field produced by the magnetization itself. As a consequence, the resonance
frequency is mainly defined by

γ
fR = JS (50)

γ
where 2π 28 GHz/T. The working frequency is usually chosen in the vicinity of
fR , higher in the upper band, lower in the lower band. So from (50) it is clear that a
low magnetization is necessary for UHF, whereas a high one is needed in W-band.
At constant frequency, the material shows an absorption peak (a maximum in μ ) as
a function of the DC field. This absorption peak is characterized by its FWHM, ΔH
expressed in A/m and is the main characteristic of microwave ferrites. Depending
on the application, the line width ΔH should be sharp or large and can be tuned
according to [28]

3 μ0 K1
ΔH = ΔHi + pJS + 2.07 3 (51)
2μ0 JS

where the first term – due to gyromagnetic damping – is intrinsic (see  Chap. 26,
“Electron Paramagnetic and Ferromagnetic Resonance”), the second term corre-
sponds to the broadening due to porosity (p), and the last one is due to spin wave
broadening (the factor 2.07 is for a sphere). Because porosity is not desirable, as
it is detrimental to the losses, ΔH is tuned by adjusting JS with substitutions. In
nonreciprocal applications, such as gyrators or isolators, the DC field is constant and
can be produced by a magnet. In reciprocal applications, a DC field is applied to the
material by a permanent magnet, and the microwave field is applied perpendicularly
(or less commonly parallel in some applications). In tunable phase shifter, the field
is of course variable and produced by an electromagnet. In applications exploiting
the Faraday effect (rotation of the polarization plane), the ferrite is not polarized.
Microwave ferrites are of two types: spinels and garnets; their main characteris-
tics are given in Table 15.
Microwave spinels are very similar to other soft ferrites except the substitution
type or rate:

– NiZn microwave ferrites contain less Zn in general (between 0 and 0.3); they
are cheapest microwave ferrites since they are massively produced for other
purposes.
– Magnesium ferrites, in which Mg is combined with Mn as divalent element,
which exhibits lower loss than NiZn ones.
– Lithium ferrite , where Li+ is combined with trivalent iron according to the
formula (Li+ 3+
0.5 Fe0.5 )1−x Znx Fe2 04 , which has sharper absorption line.
1480 F. Mazaleyrat

Table 15 Properties of selected microwave ferrites [39, 6, 30]. For gadolinium-iron garnets,
the compensation point of the magnetization is indicated with the Curie point. The dielectric
permittivity lies in the range 13–16
Composition JS (T) TC (°C) ΔH (kA/m) Frequency of use (GHz)
25°C
NiFe2 04 0.34 585 32 10–40
Ni0.6 Zn0.4 Fe2 04 0.5 350 13 10–40
Li0.5 Fe0.5 Fe2 04 0.36 640 28 10–40
Li0.3 Fe0.3 Zn0.4 Fe2 04 0.50 425 8 10–40
Y3 Fe5 012 0.19 290 6 1–10
Y1.5 Gd1.5 Fe5 012 0.096 −150/290 14 1–10
Gd3 Fe5 012 0.009 27/290 broad 1–10
Y2.4 Ca0.6 Fe4.7 V0.3 O12 0.122 235 4 1–10
Y1.4 Ca1.6 Fe4.1 V0.8 O12 0.036 225 40 1–10
Y3 Fe4.75 In0.25 O12 0.19 207 3 1–10
Y1.8 Ca1.2 Fe3.8 V0.6 In0.6 O12 0.1 109 1 1–10

Among microwave spinels, lithium one shows the best properties (small ΔH , high
TC , low loss), but they are much more difficult to produce since Li partly evaporates
during firing.
Yttrium iron garnet (YIG) has the canonical composition (Y3 Fe5 012 ) where
all metal ions are trivalent. Here, again, there are a lot of possible substitutions
providing the ion has charge and ionic radius compatible with the crystallographic
site in which it is supposed to go. Most common substitutions are:

– Al for Fe which reduced JS for VHF/UHF applications.


– Gd for Y which produces a compensation point2 of the magnetization and a
positive magnetization variation with temperature in a range tunable by the Gd
content.
– A combination of Gd and Al to combine the abovementioned effects.
– Fe by In in order to increase the magnetization (using the same effect as Zn in
spinels).
– Ca2+ is also substituted for Y when combined with V4+ to enhance the
magnetization.

YIG is the only magnetic material transparent to light (in the near infrared) and
displays strong Faraday effect ; it is thus also used in magneto-optical devices. Due
to its compensation point, Gd-YIG thermal variation of the saturation magnetization
is rather low around room temperature. Upon all microwave ferrites, YIG has the
lowest dielectric loss (tan δ = 2 · 10−4 at 8.2 GHz and ≈5 for spinels).
Microwave ferrites are produced using the same processes as spinel ferrite but
in particular shapes: sphere (in the past), disk, square, or triangle (usually with cut
corners) from 0.5 to 3 mm thick and up to 50 mm large.
29 Soft Magnetic Materials and Applications 1481

Effect of a Gap on Magnetic Properties

Cores with gap are used in every application where inductive energy storage is
necessary such as smoothing chokes or coupled inductors because most of soft
magnetic material cannot store energy (except stress annealed Finemet and in a
lower extent transverse field-annealed amorphous alloys and permalloy). Gaps can
be either lumped or spread depending on the application.

Core with Lumped Gap

Core with lumped gap is made of any kind of soft magnetic materials depending
essentially on the frequency of the device (silicon steel, Fe-Ni, Mn-Zn ferrite,
amorphous, etc.). They are exactly similar to single-phase transformer cores except
the fact stacked cores are stacked without overlapping joints and that tape wound
cores are more frequently used. Rectangular tape wound core is cut into two pieces
after impregnation (C-cores), whereas ring cores are molded in resin before cutting
a single gap. The effective permeability of the core (assuming a constant cross-
sectional area along flux lines) is expressed in a very similar way as (49):

μR
μeff = (52)
1 + μR σg


where g is the length of the gap,  is the length of the magnetic path, and σ is the
dispersion coefficient due to flux spreading in the gap. For small gaps, i.e., when
g is small compared to  and to the dimensions of the cross section, σ 1. If
this condition is not realized, μeff = σg 
and σ is usually smaller than 0.5. The
advantage of lump gap core is the possibility to tune the permeability by the user.
However, flux leakage around the gap is responsible for electromagnetic emission
and additional losses due to the nonuniformity of the induction at the gap edges.

Cores with Spread Gap

Cores with spread gap are used to overcome the drawbacks of lumped cores,
especially in the RF bands. They are realized from metallic powders mixed with
a resin and pressed. The filling factor is adjusted to the target permeability. The
effective permeability can be expressed as a function of the filling factor, φ, the
volume ratio occupied by ferromagnetic particles in the core:

μR
μeff = (53)
1 + μR (1 − φ 1/3 )

for large concentrations or for low concentrations satisfying μR (1 − φ)  3 by the


Clausius-Mossotti formula:
1482 F. Mazaleyrat

1 + 2φ
μeff = (54)
1−φ

The main advantages of powder core compared to lumped gap cores are lower
losses for a given effective permeability and the possibility to shape cores by
pressing or injection molding in the form of rings, E, L, screws, pots, cups, tubes,
rods, etc.

Iron Powder Cores


Iron powder is used for high frequency because of the large magnetization and
magnetocrystalline anisotropy which pushes the ferromagnetic resonance above
1 GHz. For frequencies above 10 MHz, particles have to be smaller than 10 μm to
limit eddy currents. Iron carbonyl particles are made by decomposition of Fe(CO)5 .
They contain impurities due to the process (O, N, C) which contributes to decrease
in the permeability. They have a grain size distribution between 1 and 10 μm and
an oxide surface layer. Due to the small grain size, the particle permeability is
small according to (47), and the filling factor is low, so the effective permeability is
maximum 20 in practice.
For higher permeabilities, different methods for pure iron particle production
can be used, but the most efficient is atomization in inert gas or rotating electrode
sputtering because these techniques give powder with good sphericity and low size
dispersion. The permeability depends on the size of the particles (usually from
several tens to about 200 μm), the binder quantity, the pressure, and the temperature
of compaction. Depending on the filling factor and the grain size, the permeability
extends from several tens to nearly 1000. Of course, as φ increases, the resistivity
drops due to contact between particles and percolation (see Table 16).
Iron powder cores are used in a great variety of applications, from power
components in RF radio emitters to HF power electronics. High-density hot pressed

Table 16 Magnetic properties of selected powder cores. φ is the filling factor; D is the mean
grain size [26, 9]
Composition μR JS B10k HC φ D ρ fmax loss mW
(T) (T) (A/m) (%) (μm) (Ωm) (Hz) cm3
Carbonyl iron 11 1.6 0.15 1600 60 4–8 5.105 80 M P125mT
MHz =550
Carbonyl iron 20 1.9 0.3 1600 70 < 10 1 60 M P125mT
MHz =800
Iron 40 1.0 0.9 1200 45 1 2M 25mT
P100 =450
kHz

Iron 75 1.8 1.3 900 62 10−2 100 k 25mT


P100 kHz =250
Iron 220 2.0 1.33 200 91 40 18.10−3 200 k P11TkHz = 28
Iron 950 2.1 1.63 200 96 100 70.10−6 20 k P11TkHz = 19
Mo-permalloy 14 0.57 0.03 400 72 60 9M 25mT
P100 = 73
kHz

Mo-permalloy 160 0.74 0.2 400 92 200 700 k 25mT


P100 = 37
kHz

Mo-permalloy 550 0.76 0.47 400 95 200 90 k 25mT


P100 =266
kHz
29 Soft Magnetic Materials and Applications 1483

powder cores, aka soft magnetic composites (SMC), exhibit permeabilities between
200 and 1000; they are used for chokes and electrical motors, respectively.

Mo-Permalloy Powder Cores


Molybdenum permalloy powders cores, aka MPP, are made by the same methods
as the pure iron ones. They have larger permeability for the same filling factor and
lower losses for the same perm eability due to the vanishing magnetocrystalline
anisotropy. On the other side, they have lower bandpass due to the smaller
JS and K1 . MPP are used for high-frequency power electronics as smoothing
chokes, resonance inductor, inverter inductors, or flyback SMPS coupled inductor
(Table 18).

Other Powder Cores


Besides the abovementioned common powder core, there are also niche markets for
more exotic materials:

– Fe50 Ni50 is a compromise between high B iron and low-loss MPP, for μR = 160,
25mT = 70 mWcm−3 .
P100 kHz
– Sendust is an iron alloy containing 9% Si 6% Al, a composition very close to
the ordered Fe3 (Si, Al), that shows characteristics quite similar to MPP (K1 and
λS 0) but with higher magnetization (up to 1 T), for μR = 160, P100 kHz =
25mT
−3
30 mWcm .
– 6.5 silicon steel is also used for its zero magnetostriction (low noise in the audio
band like MPP or Sendust) but with higher losses and lower permeability due to
nonzero K1 .

Annex 1: Comparison of Materials for Power Application in the


VLF to HF Band

Table 17 Comparison of losses in different magnetic materials used in power electronics. All
values are indicated for the same B × f = 25 kVm−2 . Losses are expressed in mW · cm−3
B (mT) 1000 250 100 50 25 10 5
f (kHz) 25 100 250 500 1000 2500 5000
3C93 1000 800 300
3F35 80 120
4F1 100 80
N27 1400 1000
N49 1000 300 120 80
N72 400
MP205 7000 6000 6000
Sendust160 4000 2700 2000
2714A 683 759
Finemet 1900 759
1484 F. Mazaleyrat

Annex 2: Materials Selection Table

Table 18 In the following table, different classes of applications are reviewed, and materials are
proposed for the purpose. These propositions are only guidelines; the grade must be adapted to the
application specifications (frequency, shape, cost, etc.). VH stands for very high, H for high, VL
for very low, and L for low
Application Required properties Material type Typical grade
Network transformer H B, VL 50 Hz loss GO Fe-Si M090-23P
HV/LV network transformer H B, VL 50 Hz loss Fe-based glass Metglas SA
H B, VL 50 Hz loss GO Fe-Si M110-27S
LV industrial transformer H B, L 50 Hz loss GO Fe-Si M110-35-N
Airborne transformer VH B Fe-Co 49%Co
Household transformer L cost L Si steel M600-50-A
Power station generator stator H B, VL loss GO Fe-Si M090-30-N
Power station generator rotor strength Forged steel 95%Fe+Ni, Cr, Mo, C
Boat/train motors H B, L loss NO FeSi M235-35-A
Industrial motors H B, L 50 Hz loss NO FeSi M330-35-A
High-speed motors H B, L VF-band loss NO FeSi M270-20-A
Automobile motors H B, VL loss NO FeSi M210-27-A
Automobile generators H B, L cost NO FeSi M400-50-A
Claw pole rotor Forgeability C steel C05E
Airborne generators VH B, Strength FeCo 25%Co
Electromagnetic motor valve VH B FeCo 49%Co
Household AC motors L cost L Si steel M600-50-A
Small DC motor stator Stampability Steel Low C
Small DC motor rotor L cost L Si steel M600-50-A
Watch step motor H B and μ Fe-Ni 50%
Inductive encoders Fe-Ni 50%
Forward SMPS transformer L 10kHz loss Co amorphous F loop
Co-based F loop
Nanocryst. F loop
MnZn ferrite Power grade
Flyback SMPS transformer BR = 0, L μ MnZn ferrite Power grade+gap
Nanocrystalline Stress annealed
Smoothing choke Linearity, f<10 kHz FeSi Wound core+gap
Linearity, f<100 kHz MnZn Gap
Linearity, f<100 kHz Powder core MPP or iron
Linearity, f<100 kHz Nanocrystalline Stress annealed
Linearity, f>100 kHz Carbonyl iron
Resonance SMPS L LF-band loss MnZn ferrite Power grade
Powder core MPP
L MF-band loss NiZn ferrite Power grade
MPP Carbonyl iron
Wideband transformer Hμ Signal grade
Drive transformers for IGBT H μ, linearity FeNi Field-annealed permalloy
(continued)
29 Soft Magnetic Materials and Applications 1485

Table 18 (continued)
Drive transformer for GTO H BR , H dynamics FeNi Textured 50%Ni
Network inductor HB GO FeSi N-grade wound core +
gap
Pulse transformer H BR , H dyn. Co amorphous Z loop
Spike killer H BR , H dyn. Co amorphous Z loop
Delay lines H μ, bar shape MnZn ferrite Signal grade
Magnetic amplifier H BR GO FeSi N-grade wound core
Current transformer Hμ Permalloy F loop
Co amorphous F loop
Nanocryst. F loop
Thick-layer micro-inductor Low sintering temp. Ni-Zn-Cu ferrite Tape casting
On-ship inductance Integration YIG RF sputtered on
substrate
On-ship transformer Hμ Fe-Ni Electrodeposited
Flux-gate sensor VH BR , VL HC Co amorphous F loop
Hall effect sensors L BR , L HC MnZn ferrite Signal grade
Magnetostrictive sensor HλS Amorphous Fe/FeNi based
Proximity sensors H MF-band μ MnZn ferrite
Ground fault circuit breaker VH μ Co amorphous F loop
Nanocryst. F loop
FeNi Permalloy
Common mode filter VH μ MnZn ferrite Signal grade
Nanocrystalline Field annealed
LF filter Lμ MnZn ferrite Signal grade + gap
MF-UHF filter Lμ NiZn ferrite
Telecom transformer Hμ MnZn ferrite Signal grade
LF antennae Linearity, high μ MnZn ferrite
MF antennae Bandwidth NiZn ferrite μ < 1000
HF antennae Bandwidth NiZn ferrite μ < 250
VHF antennae Bandwidth NiZn ferrite μ < 100
UHF antennae Bandwidth NiZn μ < 20
Radar absorber UHF absorption NiZn Plate, structured surface
Faraday effect uniline HC < 100A/m Li ferrite Rod
Resonator filter L loss, L ΔH (Ga)YIG Single crystal sphere
Circulator, phase shifter f < 1.5 GHz Gd-YIG Plate, rod
1.5 < f < 10 GHz Al-YIG id.
f > 10 GHz Li or NiZn spinel id.
Wireless energy transfer H B, L MF-band loss Ferrite MnZn
Zero Tesla chamber Hμ FeNi mumetal
MF magnetic shielding MnZn ferrite Signal grade
Flexible magnetic shielding H μ Nanocrystalline film-coated ribbon
Large bandwidth MnZn ferrite Powder in polymer film
Induction cooking pan Controlled TC FeNi Thermal alloy
DC electromagnet HB Fe/low C steel Bulk
AC electromagnet HB FeSi M600-100-A
1486 F. Mazaleyrat

Notes
1 Bulk metallic glasses have been successfully produced by the end of the 1990s, but they have

relatively low magnetization. In addition, the bulk state is a drawback for applications, so they
have never been produced at industrial scale.
2 The compensation point, see  Chap. 17, “Magnetic Oxides and Other Compounds.”

References
1. Ahmadi, B., Zehani, K., LoBue, M., Loyau, V., Mazaleyrat, F.: Temperature dependence of
magnetic behaviour in very fine grained, spark plasma sintered NiCuZn ferrites. J. Appl. Phys.
111(7), 07A510 (2012)
2. Alben, R., Becker, J.J., Chi, M.C.: Random anisotropy in amorphous ferromagnets. J. Appl.
Phys. 49, 1643 (1978)
3. Alves, F., Simon, F., Kane, S.N., Mazaleyrat, F., Waeckerle, T. Save, T., Gupta, A.: Influence
of rapid stress annealing on magnetic and structural properties of nanocrystalline Fe74.5 Cu1
Nb3 Si15.5 B6 alloy. J. Magn. Magn. Mater. 294(2), e141–e144 (2005)
4. J-B; Hérission, D., Benchabi, A., Waekerlé, T., Fraisse, H., Boulogne, B.A., Desmoulin,
F.: Nanocrystalline material strip production for the formation of magnetic torus comprises
annealing cast amorphous ribbon of specified composition defiling under tension (2002)
5. Amperam.: Nanocrystalline Core Solution (2012)
6. Anderson, E.E., Cunningham, J.R. Jr, McDuffie, G.E.: Magnetic properties of the mixed
garnets (3-x) Y2 O3·x Gd2 O3·5 Fe2 O3 . Phys. Rev. 116(3), 624 (1959)
7. Appino, C., Khan, M., de la Barrière, O., Ragusa, C., Fiorillo, F.: Alternating and rotational
losses up to magnetic saturation in non-oriented steel sheets. IEEE Trans. Magn. 52(5), 1–4
(2016)
8. Barbisio, E., Fiorillo, F., Ragusa, C.: Predicting loss in magnetic steels under arbitrary
induction waveform and with minor hysteresis loops. IEEE Trans. Magn. 40(4), 1810–1819
(2004)
9. BASF.: Carbonyl Iron Powder for Inductive Electronic Components
10. Beatrice, C., Appino, C., de la Barrière, O., Fiorillo, F., Ragusa, C.: Broadband magnetic losses
in fe-si and fe-co laminations. IEEE Trans. Magn. 50(4), 1–4 (2014)
11. Bertotti, G.: Hysteresi in Magnetism. Academic Press, San Diego (1998)
12. Bertotti, G.: General properties of power losses in soft ferromagnetic materials. IEEE Trans.
Magn. 24(1), 621–630 (1988)
13. Bertotti, G.: Dynamic generalization of the scalar preisach model of hysteresis. IEEE Trans.
Magn. 28(5), 2599–2601 (1992)
14. Bertotti, G., Mayergoyz, I.D. (eds).: The Science of Hysteresis: Physical Modeling, Micromag-
netics, and Magnetization Dynamics. Academic Press, Oxford (2006)
15. Brailsford, F.: Physical Principles of Magnetism. van Nostrand, London (1966)
16. Cazin, A.: Sur les effets thermiques du magnétisme. J. Phys. Theor. Appl. 5(1), 111–118 (1876)
17. de la Barriere, O., Ragusa, C., Appino, C., Fiorillo, F., LoBue, M., Mazaleyrat, F.: A
computationally effective dynamic hysteresis model taking into account skin effect in magnetic
laminations. Phys. B Condens. Matter 435, 80–83 (2014)
18. Ewing, J.A.: Induction in iron and other metals. The Electrician, 3rd edn. The Electrician
Publishing co, London (1901)
19. Glaser, A.A., Kleynerman, N.M., Lukshina, V.A., Patapov, A.P., Serikov, V.V.: Thermome-
chanical treatment of the nanocrystalline alloy fecunbsib. Phys. Met. Metal. 72, 53 (1991)
20. Herzer, G.: Grain structure and magnetism of nanocrystalline ferromagnets. IEEE Trans. Magn.
25, 3327–3329 (1989)
21. Herzer, G.: Grain size dependence of coercivity and permeability in nanocrystalline ferromag-
nets. IEEE Trans. Magn. 26, 1397–1402 (1990)
29 Soft Magnetic Materials and Applications 1487

22. Hitachi Metals: Metglas amorphous alloys. https://Metglas.com


23. Hitachi Metals: Nanocrystalline soft magnetic material FINEMET (2006). http://www.hitachi-
metals.co.jp
24. JFE Steel Corporation: JFE Super Core (Electrical steel sheets for high-frequency application)
(2005)
25. Lebourgeois, R., Duguey, S., Ganne, J-P., Heintz, J-M.: Influence of v 2 o 5 on the magnetic
properties of nickel–zinc–copper ferrites. J. Magn. Magn. Mater. 312(2), 328–330 (2007)
26. Micrometals: Arnold powder cores. https://micrometals.com/, https://www.mag-inc.com/
27. Orb Electrical Steels: Cogent™Non oriented electrical steel, typical data (2015)
28. Patton, C.: A review of microwave relaxation in polycrystalline ferrites. IEEE Trans. Magn.
8(3), 433–439 (1972)
29. Pry, R.H., Bean, C.P.: Calculation of the energy loss in magnetic sheet materials using a domain
model. J. Appl. Phys. 29(3), 532–533 (1958)
30. Sirvetz, M.H., Zneimer, J.E.: Microwave properties of polycrystalline rare earth garnets. J.
Appl. Phys. 29(3), 431–433 (1958)
31. Lucas, A., Lebourgeois, R., Mazaleyrat, F., Laboure, E.: Temperature dependence of spin
resonance in cobalt substituted NiZnCu ferrites. Appl. Phys. Lett. 97(18), 182502 (2010)
32. Lucas, A., Lebourgeois, R., Mazaleyrat, F., Laboure, E.: Temperature dependence of core loss
in cobalt substituted Ni-Zn-Cu ferrites. J. Manetism Mag. Mater. 323(6), 735–739 (2011)
33. Mazaleyrat, F. Varga, L.K.: Thermo-magnetic transitions in two-phase nanostructured materi-
als. IEEE Trans. Magnetics Trans. Magnet. 37(4), 2232–2235 (2001)
34. ThyssenKrupp Electrical Steel GmbH: Grain oriented electrical steel PowerCore® (2012)
35. Toda, H., Oda, Y., Kohno, M., Ishida, M., Zaizen, Y.: A new high flux density non-oriented
electrical steel sheet and its motor performance. IEEE Trans. Magn. 48(11), 3060–3063 (2012)
36. Tsutaoka, T.: Frequency dispersion of complex permeability in mn–zn and ni–zn spinel ferrites
and their composite materials. J. Appl. Phys. 93(5), 2789–2796 (2003)
37. Vacuumschmelze: Soft Magnetic Materials and Semi-finished Products (2005)
38. Warburg, E.: Magnetische untersuchungen. Annalen der Physik 249(5), 141–164 (1881)
39. WEST, R.G., Blankenship, A.C.: Magnetic properties of dense lithium ferrites. J. Am. Ceram.
Soc. 50(7), 343–349 (1967)
40. Yoshizawa, Y., Yamaguchi, K.: New Fe-based soft magnetic alloys composed of ultrafine grain
structure. J. Appl. Phys. 64, 6044 (1988)

Frédéric Mazaleyrat graduated in electrical engineering in 1993


and received his PhD in 1996 from Université Pierre & Marie
Curie. (now Sorbonne Université). He worked first on nanocrys-
talline soft magnetic materials and then on the development of
soft and hard magnetic materials, synthesis, structural dependence
of magnetic properties, modeling of functional properties and
applications. He’s presently a full professor at École normale
supérieure, Université Paris-Saclay, teaching electrical engineer-
ing and applied physics. FM is also an editor for the IEEE
Transactions on Magnetics (since 2020), the Intermag confer-
ence (since 2019 and the French encyclopedia Techniques de
l’Ingénieur (since 2000).
He has authored more than 130 papers in peer reviewed jour-
nals on soft and hard magnetic materials and their applications.
Magnetocaloric Materials and Applications
30
Karl G. Sandeman and So Takei

Contents
Introduction: The Magnetocaloric Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1490
Cooling at Low or Intermediate Temperatures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1494
A Brief History of the Magnetocaloric Effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1494
Adiabatic Demagnetization (ADR) Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1495
Nuclear Adiabatic Demagnetization (NDR) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1497
The Magnetocaloric Effect Near Room Temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1498
Room Temperature Magnetocaloric Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1498
Room Temperature Magnetic Cooling Engines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1509
Power Conversion and Spin Caloritronics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1511
Power Conversion via Pyromagnetic Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1511
Spin Seebeck and Spin Nernst Effects and Spin Caloritronics . . . . . . . . . . . . . . . . . . . . . . 1513
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1518

Abstract

The physics of magnetocaloric, pyromagnetic, and spin Seebeck effects in


magnetic materials is presented. All are magnetothermal effects that can be used
in a range of cooling, power generation, or spintronic devices. The focus is on
the materials that have generated the most interest for research and applications,
beginning with cooling at millikelvin and microkelvin temperatures via the

K. G. Sandeman ()
Department of Physics, Brooklyn College of the City University of New York, Brooklyn,
NY, USA
The Physics Program, The Graduate Center, CUNY, New York, NY, USA
e-mail: karlsandeman@brooklyn.cuny.edu
S. Takei
The Physics Program, The Graduate Center, CUNY, New York, NY, USA
Department of Physics, Queens College of the City University of New York, Flushing, NY, USA
e-mail: so.takei@qc.cuny.edu

© Springer Nature Switzerland AG 2021 1489


J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_13
1490 K. G. Sandeman and S. Takei

magnetocaloric effect. Then, the range of materials, often with first-order phase
transitions, that are the feature of current magnetocaloric effect studies near room
temperature is presented. Lastly, progress is examined in utilizing the reverse
magnetocaloric (pyromagnetic) effect for power generation and in using the
spin Seebeck, spin Hall, and spin Nernst effects in thin-film heterostructures for
spintronics and caloritronics.

Introduction: The Magnetocaloric Effect

The magnetocaloric effect (MCE) is the change of temperature of a material that


results from a change in applied magnetic field. The simplest explanation of the
effect is perhaps in terms of a picture of localized magnetic moments; when such
moments are aligned by a magnetic field, the usual consequence is a reduction of
their entropy. If performed adiabatically, the application of the magnetic field also
results in an opposing increase of other entropies, such as the entropy of vibrations
of the crystal lattice. In that case, the adiabatic application of a magnetic field
causes heating, while adiabatic removal of a magnetic field causes cooling. This
scenario is known as the conventional magnetocaloric effect. From the point of view
of total free energy, the magnetic field normally stabilizes a high magnetization
state that exists at lower temperature than the initial temperature of the system.
Alternatively, the magnetic field can, in some cases, stabilize a high magnetization
state that exists at a temperature higher than the starting temperature. In this
case, the thermodynamics described above is reversed, and a so-called inverse
magnetocaloric effect results, in which a material cools when a magnetic field is
applied adiabatically.
A thorough review of the thermodynamic principles of the magnetocaloric effect
can be found in references [1] (for low temperatures) and [2] (for near room
temperature). We may use equilibrium thermodynamics to describe many examples
of magnetocaloric effects in which there is continuity of the (magnetic) order
parameter throughout the range of temperature and magnetic field studied. The
starting point is the free energy:


dG = −SdT − MdH + Ki dσi + . . . , (1)
i

where the symbols here have their usual meanings: S is (total) entropy; M is the
magnetisation; Ki are the anisotropy constants, and σ i is the relevant direction
cosines for the crystal symmetry of the particular material. If we consider variations
in the magnitude (rather than direction) of the magnetization, then from this free
energy, a Maxwell relation that connects the thermal variation of magnetization to
the field dependence of total entropy may be written:
30 Magnetocaloric Materials and Applications 1491

   
∂S ∂M
= . (2)
∂H T ∂T H

This Maxwell relation may be integrated to obtain the change of total entropy
that results from the isothermal application of a magnetic field:

 
 Happlied  ∂M 
S T , Happlied = dH. (3)
0 ∂T H

Equation (3) is the most often used thermodynamic relation in magnetocaloric


materials research since magnetization data can be relatively easily obtained and
can be used to estimate isothermal entropy changes through application of Eq. (3).
However, the aforementioned adiabatic temperature change is an equally important
material parameter. It may be obtained through a combination of measurements of
field- and temperature-dependent heat capacity, c(T, H), and of magnetization:

  
  Happlied T ∂M
Tad T , Happlied = − dH (4)
0 c (T , H ) ∂T H

Alternatively, the adiabatic temperature change may be measured directly by


contact or non-contact thermometry. For a fuller description of measurement
methods for both s and Tad , see [3].
From Eqs. (3) and (4), we see that the sign of both S and Tad is determined by
the sign of ∂M
∂T and thence by the way in which magnetization varies with temper-
ature. A conventional magnetocaloric effectis associated with a magnetization that
falls with increasing temperature ∂M ∂T < 0 , while inverse magnetocaloric effects
∂M
require positive values of ∂T . Relations equivalent to Eqs. (3) and (4) also exist for
the temperature variation of the anisotropy constants; in this case, the isothermal
entropy change and adiabatic temperature change result from a change in the angle,
rather than magnitude, of the applied magnetic field. The magnetocaloric effect
in that case is sometimes known as the rotational MCE and can in principle also
be conventional or inverse [4]. The relationship between the magnetization, S
and Tad , is shown in the case of a conventional MCE material in Fig. 1. The
figure also shows a feature of some materials, namely, the ability to tune from a
sharp, first-order magnetic transition through a tricritical point to a second-order
transition.
By extension, the magnetocaloric effect has parallels in electrically polarizable
materials (the electrocaloric effect) and in those in which the total entropy can be
influenced by an applied stress or pressure (leading to the mechanocaloric effect).
To see this, we may include relevant terms in the free energy in Eq. (1):
1492 K. G. Sandeman and S. Takei

Fig. 1 A schematic of the conventional MCE associated with a material that can possess a
sharp first-order or gradual second-order phase transition with ∂M
∂T < 0, depending on material
composition or magnetic field. The first-order region is associated with larger magnetization
gradient and hence larger S and Tad . (Aspects of this figure have been reproduced from I.
Takeuchi and K.G. Sandeman, Phys. Today 68, 48 (2015) [5] with the permission of the American
Institute of Physics. An inverse magnetocaloric material possesses the opposite sign of ∂M
∂T , S,
and Tad )


dG = −SdT − MdH + Ki dσi + V dp − P d + . . . , (5)
i

By analogy with the above analysis, we can see that the electrocaloric effect,
the temperature change induced by a change in the magnitude of an electric field,
 applied , is
  
  Eapplied T ∂P
Tad T , Eapplied = − dE. (6)
0 c (T , E) ∂T 

Similarly, the barocaloric effect, the temperature change induced by a change in


an applied pressure, p, is
  
  papplied T ∂V
Tad T , papplied = dp. (7)
0 c (T , p) ∂T p

The last of these relations is in fact key to conventional, gas compression cooling.
The temperature change of a material induced by a change in applied pressure  is
greatest when there is a large variation of material volume with temperature ∂V ∂T .
For this reason, a material with a liquid-gas phase transition is an ideal candidate for
observation of a large barocaloric effects, and such materials have been the majority
of (gas-compression) refrigerants to date. In gas-based cooling devices, a volatile
liquid is expanded through a nozzle and cools (Fig. 2). It then exchanges heat with
the items to be cooled, before being compressed. The compression stage raises the
30 Magnetocaloric Materials and Applications 1493

Fig. 2 A schematic comparison of a gas-based cooling cycle (a) and a solid-state ferroic cooling
cycle (b). (These figures have been reproduced from I. Takeuchi and K.G. Sandeman, Phys. Today
68, 48 (2015) [5] with the permission of the American Institute of Physics. We note that, in the
latter case, a separate fluid is needed to exchange heat with the ferroic refrigerant)

refrigerant’s temperature, and so the last stage in the cooling cycle is an exchange
of heat with the surroundings (e.g., via a radiator).
At this point, it is instructive to take stock of the basic thermodynamics presented
above to distinguish two means of providing (magneto)caloric effects. The first is
in when the applied field does not induce a phase transition; then the values of
Tad (and S) will be moderate or small. Exceptions occur when the temperature
is low enough either for the phononic part of the heat capacity to be quenched
or for the magnetic susceptibility – especially in the case of a paramagnet – to
become large. Stated in another way, without a phase transition, the entropy change
associated with magnetic moment rotation (the paraprocess) is small compared
to the entropy required to change the ordering of phonons, especially at elevated
temperatures. As a result, this version of the magnetocaloric effect has been used
in low temperature research applications and has become known as adiabatic
demagnetization refrigeration (ADR; see Sect. “Adiabatic Demagnetization (ADR)
Materials”), although the process of adiabatic demagnetization can apply to a
cooling cycle with or without a phase transition in the refrigerant.
The second type of magnetocaloric effectarises  around a phase transition. In that
case, larger (potentially infinite) values of ∂M ∂T H can be found, and – especially
in the case of a first-order phase transition – large values of Tad and S can be
observed, not just at low temperatures but also at elevated temperatures. In this case,
the applied field can induce a large (phase) change in the entropy of the magnetic
subsystem and thereby a large change in the phononic subsystem. Some numerical
values for both Tad and S around room temperature will be provided for the
materials discussed in the following sections.
The surge in interest in magnetic cooling during the last 20 years has stemmed
primarily from interest in the latter examples of MCEs observed in first-order
transition materials, motivated by the potential use of magnetic refrigerants in the
room temperature range as replacement working materials in a gas-free cooling
1494 K. G. Sandeman and S. Takei

engine. There are two main drivers. Firstly, many of the refrigerants used around
room temperature in gas compression systems are greenhouse gases. The volatility
that is vital to their operation therefore poses a risk of environmental damage
if a leak or improper recycling occurs. Some predictions put the fractional CO2
equivalent contribution of refrigerant and aerosol gases as high as 45% by 2045
[6], while the choice of alternative gas-based refrigerants is somewhat restricted
[7]. Secondly, improvements to the efficiency of modern refrigerators are costly
and involve the use of variable speed compressors that modulate the delivered
cooling power depending on demand. Cost-effective improvements of efficiency
from magnetic (or any other gas-free) cooling method are welcome. The relevance
of this statement is heightened by the large percentage of energy (at least 15% in
general) that is used for cooling, depending on the country, or type of cooling (air
conditioning, refrigeration, and so on) [8].
Before discussing examples of magnetic refrigerants and magnetic cooling in
different temperature ranges, we will first recap the history of the magnetocaloric
effect (Sect. “Cooling at Low or Intermediate Temperatures”), including a short
discussion of low-temperature cooling using electronic paramagnets and nuclear
magnetic moments. Then, in Sect. “The Magnetocaloric Effect Near Room Tem-
perature” we will focus on room temperature effects. Whichever mechanism of the
MCE (paraprocess, rotation, etc.), the search for ever larger MCEs around room
temperature remains the motivation for much of the current materials research in the
field. Since this chapter focuses on caloric material properties, we refer the reader to
references [9,10,176] for a discussion of the status of magnetic cooling prototypes
and magnetic field sources used in such devices. Section “Power Conversion and
Spin Caloritronics” contains a brief survey of two materials research fields: power
generation using pyromagnetic effects (the corollary of the magnetocaloric effect)
and thin-film caloritronics, where a thermal gradient can be used to generate a spin
current.

Cooling at Low or Intermediate Temperatures

A Brief History of the Magnetocaloric Effect

The history of the magnetocaloric effect has recently been re-examined by Anders
Smith [11]. While theoretical studies may be traced to Lord Kelvin’s prediction
that iron in its ferromagnetic state will cool when demagnetized [12], experimental
work began with Pierre Weiss and Auguste Piccard’s observations, in 1917, of a
magnetic field-induced, reversible, heating effect in ferromagnetic nickel around its
Curie temperature (354 ◦ C) [13]. In what follows, we will see that Kelvin, Weiss,
and Piccard’s works are important foundations for room temperature magnetic
cooling.
Quite separately, in 1905 Paul Langevin had predicted the existence of a
temperature change of a paramagnet when exposed to a magnetic field. It was
Langevin’s work that seems to have inspired theoretical comments on adiabatic
demagnetization at low temperature by Debye in 1926 [14] and by Giauque in 1927
30 Magnetocaloric Materials and Applications 1495

[15]. The first experimental studies of cooling by removal of a magnetic field from
paramagnetic salts at low temperature were reported by Giauque and MacDougall
in 1933 [16] and soon after by De Haas, Wiersma, and Kramers [17]. Giauque and
MacDougall were able to reach a temperature of 0.53 K when a magnetic field
of 0.8 Tesla (T) was removed from a sample of Gd2 (SO4 )3 ·8H2 O at a starting
temperature of 3.4 K. When starting from 1.5 K, they were able to reach 0.25 K and
could infer that multiple demagnetizations would result in yet lower temperatures.
From this brief history, we may see that the use of magnetic cooling at different
temperatures evolved somewhat independently. Giauque received the 1949 Nobel
Prize in Chemistry for his work on the thermodynamic properties of substances at
low temperatures, and particular mention during the presentation speech was made
of his work on magnetic cooling and the subsequent low temperature research that
had been enabled by it [18]. The field of higher-temperature magnetic cooling,
using a refrigerant at a temperature near its magnetic phase transition, received
sporadic interest after Weiss and Piccard’s experiments and, until the 1990s,
lacked the fundamental or commercial motivation for a large effort by the research
community. Proposals to use magnetic phase transitions in the opposite process,
namely, conversion of heat to power, were also sporadic and arguably predate
magnetic cooling investigations. They are described in Sect. “Power Conversion
and Spin Caloritronics”.

Adiabatic Demagnetization (ADR) Materials

An introduction to ADR materials and their uses in a cooling cycle in the millikelvin
range is given by Wikus et al. [1]. As the authors detail, a range of paramagnetic
materials can be used for different temperatures, ranging from GdLiF4 (minimum
operating temperature 500 mK) and GGG (minimum temperature 200 mK) down
to Ce2 Mg3 · (NO3 )12 · 24H2 O at temperatures as low as 0.6 mK. In each case, the
basic principle is the same: the (para)magnetic dipoles are aligned by the applied
magnetic field, causing a change of entropy that can be predicted from Boltzmann
thermodynamics:
  
sinh(xy)
S = R x (coth x − y coth(xy)) + ln , (8)
sinh x

where R is the molar gas constant, x = μkBBgB T , and y = 2J + 1. The symbol g is the
Landé g-factor, and J is the total angular momentum of the dipole.
The ADR effect is of most use below a characteristic temperature where the
magnetic component of the total entropy saturates or is close to saturation, Tsat .
This is because other entropies, principally the lattice entropy, begin to dominate at
higher temperatures , and so the action of a magnetic field is less pronounced. Below
Tsat , the change of (total) entropy of the refrigerant material observed by applying
a magnetic field is sufficient to allow for heat exchange (Tad is large) and for heat
flow (S is large). A schematic ADR cycle is shown in Fig. 3.
1496 K. G. Sandeman and S. Takei

2.00

B=0
(1)

1.50
(4)
S/R [1]

1.00
(3) (2)

0.50

B=2T

0.00
0.01 0.1 1 10
T [K]

Fig. 3 Adiabatic demagnetization cycle, employing the entropy curves of a paramagnet at two
different fields, 0 and 2 T. Here, the starting temperature is 2 K (step 1), and the refrigerant finishes
the 4-step process at 100 mK (step 4). (This figure is reprinted from Cryogenics, 51, P.Wikus et al.,
“Optimum operating regimes of common paramagnetic refrigerants”, 555–558, Copyright 2011,
with permission from Elsevier [1])

Temperatures in the millikelvin range can be achieved by isothermal magnetiza-


tion and adiabatic demagnetization steps. Since the entropy of an ideal paramagnet
is a function of the ratio of applied field to the temperature (Eq. 8), we should be
able to reduce the temperature of a magnetic refrigerant at will, down to the smallest
(internal) field that acts on the magnetic moments. If the applied field is H and the
internal field is h, then in the limit that H  h, the lowest temperature achievable
for a starting temperature of Ti is [19]

Ti
Tf = h . (9)
H

For electronic paramagnets, the internal field is governed by that of the magnetic
dipole. As a result, we can expect to reach temperatures of the order of 1 mK.
Current ADR research continues to explore new materials, especially superparam-
agnetic clusters [20] and other molecular magnets [21], as potential low-temperature
ADR refrigerants.
In addition to magnetic cooling using paramagnetic salts, there are two other
topics in which research into magnetic cooling away from room temperature has
been extensive: studies of hydrogen liquefaction at intermediate temperatures using
magnetic cooling and adiabatic demagnetization using nuclear moments for ultra-
low temperature research. Regarding the former, we point the reader to a review of
30 Magnetocaloric Materials and Applications 1497

magnetically driven hydrogen liquefaction in [22,177]. The latter is described in the


next section.

Nuclear Adiabatic Demagnetization (NDR)

In Sect. “The Magnetocaloric Effect Near Room Temperature” we will examine


the cooling effect that can be obtained near room temperature by application and
removal of an applied magnetic field from an electronic solid. However, a survey
of magnetic cooling at low temperature would be incomplete without an exposition
of the technique of nuclear adiabatic demagnetization (NDR). This technique has
employed the variation in the ordering of nuclear spins to allow scientists to access
microkelvin temperatures. A concise review of the area is presented by Pickett [23],
while a longer survey of both experimental techniques and the theory of nuclear
magnetic order is given by Oja and Lounasmaa [24]. For a historical overview,
Kurti’s own recollections [19]. and those of Bleaney and Lounasmaa [25] are
recommended. Space in this chapter only permits a survey of key components of
the NDR cooling technique and of the temperatures accessed thus far.
A key bridge between a discussion of electronic magnetic cooling (ADR) and
NDR is the thermal bath in which either spin system exists. In the case of ADR,
the phonon bath exchanges heat with the electronic magnetic moments and thus
determines the temperature of the material. In the nuclear case, the electrons act as
the thermal bath. To achieve sub-millikelvin temperatures with ADR is impractical
due to the size of the dipole fields associated with electronic magnetic moments,
as outlined in the previous section. However, nuclear spins have moments that are
around 1000 times smaller, and thus microkelvin temperatures are achievable [19],
provided appropriate choices of refrigerant are made.
The idea of NDR was introduced, independently, by Gorter [26] and by Kurti
and Simon [27]. It was not until 1949 that the first experimental realization of the
technique was made using the 19 F nuclei in naturally occurring, colorless fluorite.
An applied field was lowered from 0.4 T to 0.04 T, and the nuclear temperature, as
measured by nuclear magnetic resonance, fell from 1.2 K to 170 mK. Colorless
fluorite was chosen for the experiment over naturally colored samples since the
nuclear spin-lattice relaxation time, T1 , of the former was longer (of the order of
1 s). Indeed, equilibration of the nuclei with the electrons is governed by T1 , and
so large T1 values allow more easy manipulation of the nuclear magnetic state. In
addition, a nuclear refrigerant material should not undergo any magnetic ordering
of the nuclear subsystem (which therefore prohibits materials that have magnetic
ordering associated with the electrons) and should have both a large nuclear density
of states (nuclear Curie constant) and a large density of states at the Fermi energy
to maximize coupling to the electronic subsystem [23]. From these requirements,
copper has emerged as the most used ADR refrigerant, while platinum is often used
as the nuclear subsystem thermometer on account of its high density of states at
the Fermi energy (and therefore rapid thermal response). Copper wire was first
1498 K. G. Sandeman and S. Takei

employed in an NDR system by Kurti and co-workers in 1956 using a maximum


field of 2.8 T [28].
Current NDR uses a helium dilution refrigeration state to achieve millikelvin
starting temperatures. Superconducting magnets are required to generate changes in
magnetic fields that will yield sizeable changes in the entropy of the weak nuclear
magnetic moments [24]. The rate of demagnetization is kept to below 1 T/h to
limit eddy current heating, and rods, sheets/plates, or slotted blocks of copper are
used as the refrigerant as they optimize the balance of refrigerant volume (cooling
capacity) and eddy current loss [29]. Multi-stage cascade refrigerators have been
used to access the lowest temperatures, with nuclear temperatures reaching as low
as 100 pK [30].

The Magnetocaloric Effect Near Room Temperature

Room Temperature Magnetocaloric Materials

After completing a brief survey of low temperature MCE materials research, we now
focus on the use of the magnetocaloric effect for cooling around room temperature.
This has been the subject of the majority of MCE studies, especially since the mid-
1990s.
From Eq. (4), we see that any magnetic alternative to gas-based cooling around
room temperature must have a large value of ∂M/∂T, which is most readily found
when a change of magnetic state can be invoked by the applied field. The search for
magnetic refrigerants near room temperature therefore becomes a search for phase
transitions in this temperature range. However, there are a few more conditions
that come with a room temperature magnetic refrigerant search. A large value
of ∂M/∂T is more likely when the saturation magnetization is high. Similarly,
ferro- and ferrimagnets are more viable candidates than most antiferromagnets,
since magnetization can be induced by a field around the Curie point, whereas
the appearance of a finite magnetization around a Néel point depends strongly on
the strength of the antiferromagnetic coupling. Due to the reduced magnitude of
∂M/∂T found in superparamagnets (SPMs), the MCE in that class of compounds is
generally lower at room temperature than that of a ferromagnet near its Curie point.
For further information on the MCE in SPMs, the reader is referred to [31, 32].
In terms of room temperature application, further desirable properties include
low heat capacity (to maximize Tad , see Eq. (4)); moderate thermal conductivity
(to minimize thermal equalization while still allowing rapid heat transfer to occur);
corrosion resistance (since a heat exchange fluid is required); shapeability; low
cost; toxicity; and zero critical element content. A further requirement is low
magnetothermal hysteresis (to minimize energy loss upon cycling) unless, as has
been proposed, a second driving parameter such as applied stress is used (see
later [33]). These requirements restrict the number of materials that are currently
considered as room temperature magnetic refrigerants to two principal families:
30 Magnetocaloric Materials and Applications 1499

La(Fe,Si)13 Hy and (Mn,Fe)2 (P,Si). Before discussing them in more detail, we


describe the physics of two rare earth refrigerants that have been foundational to
room temperature magnetic cooling research: Gd and the Gd-Si-Ge family.
Gadolinium metal satisfies several of the aforementioned criteria: it has a Curie
temperature of 294 K and a large saturation magnetization of 7 μB per atom and
is a single-element material, meaning that its heat capacity per volume or per unit
mass is low and so from Eq. (4) Tad is high for a material with a continuous
phase transition. In Fig. 4a, Tad data for Gd are shown for applied field changes of
0–1.1 T and 0–7 T (both upon field application). We immediately see that the
“single-shot” caloric effect that can be achieved is of the order of 3 K in a field
change of 0–1.1 T. This response exists over a range that is determined by the
gradual nature of the phase transition, with values of at least 1 K in this field
change being achievable over a temperature window of around 30 degrees. However,
unlike the ADR or NDR approaches mentioned above, the application of magnetic
cooling at room temperature is predicated on maximizing efficiency, which is
only achievable if the temperature span of the MCE is used appropriately. As
explained below, rather than employing a single-shot approach, where all parts of
the refrigerant are heated or cooled equally, a regenerator enables different parts
of the refrigerant structure (either a single material or a multi-material structure) to
operate thermodynamic cycles around different temperatures, so as to maximize the
temperature span of the final device.

14 (a) 8.6 (b)


8.4
12 0-7 Tesla
8.2 B=0 B=7 Tesla
10
Tad
Entropy/R
Tad (K)

8.0
8
S
7.8
6
7.6
4 0-1.1 Tesla
7.4
2
7.2
0
150 200 250 300 350 250 260 270 280 290 300 310 320 330
Temperature (K) Temperature (K)

Fig. 4 (a) Adiabatic temperature change data for Gd taken in a field of 0–1.1 T [35] and 0–7 T
[34]. The 0–1.1 T data have been reproduced from C. R. H. Bahl and K. K. Nielsen, J. Appl. Phys.
105, 013916 (2009), and the 0–7 T data have been reproduced from G. V. Brown, J. Appl. Phys.
47, 3673 (1976), both with the permission of the American Institute of Physics; (b) Isofield total
entropy curves in zero applied field and in 7 T. (This figure has been reproduced from G. V. Brown,
J. Appl. Phys. 47, 3673 (1976) with the permission of the American Institute of Physics)
1500 K. G. Sandeman and S. Takei

For the testing device designs, Gd has been the benchmark magnetic refrigerant
for over 40 years as, by using a single compound, a temperature span of several
tens of degrees can be achieved, depending on the applied field. In this respect,
pioneering experimental work on the use of Gd in a magnetic cooling cycle was
performed by Gerald Brown at NASA’s Lewis Research Center in 1976 [34]. In
particular, Brown mapped out the entropy vs temperature diagram of the material
from heat capacity measurements and made direct observations of the adiabatic
temperature change in a 7 T field. Brown’s S(T) data are shown in Fig. 4b together
with examples of how S and Tad relate to the S(T, B = 0) and S(T, B = 7 T)
curves. (The 0–7 T Tad data in Fig. 4a were used by Brown to construct the
S(T, B = 7 T) curve in Fig. 4b from the S(T, B = 0) curve.)
There are several important features in Fig. 4 and in the conclusions that Brown
drew. Firstly, we see that there is a large, reversible Tad response associated
with the Curie transition in Gd. As may be expected, the peak in Tad is to be
found at the Curie temperature. The S(T) data reveal a change of total entropy of
around 0.2 R per mole, where R is the molar gas constant. Brown concluded that a
Carnot cycle would not efficiently use the available change in magnetic subsystem
entropy. This is because of the presence of the phonon entropy term. From his work,
Brown concluded that an efficient use of the magnetic entropy for such a phase
transition magnet would be a magnetic Stirling cycle, involving a constant field
step rather than a constant entropy step. Brown constructed a simple regenerator
using the magnetic Stirling cycle, demonstrating that 47 K temperature span could
be developed between the hot and cold ends of a regenerator system of 1-mm-thick
Gd plates. The regenerator comprised a heat exchange fluid, which cooled or heated
the Gd in between magnetization or demagnetization steps, respectively. This kind
of regeneration had previously been studied at low temperature by van Geuns in
cycles employing paramagnets [36, 37] and would be pursued further as “active
magnetic regeneration” (AMR) by Steyert and Barclay in later studies [38,39]. The
term “active” in AMR refers to the dual function of the magnetic refrigerant as (a)
the source of the magnetocaloric effect and (b) the medium via which a thermal
gradient is established, through heat exchange with a separate working fluid [176].
As highlighted above, given that the adiabatic temperature change is of the order of a
few kelvin but that it can exist over several tens of kelvin, a regenerator is necessary
to maximize the temperature span of the cooling device. Nowadays, almost all
prototype room temperature magnetic coolers employ a regenerator, as a means of
turning a modest single-shot temperature effect into a viable thermodynamic cycle
operating between hot and cold ends of the device that are separated in temperature
by more than the single-shot Tad .
If we again examine the peak-shaped response curves shown in Fig. 4a, we
may also conclude that, in order to optimize the cooling power delivered at a
given temperature, it is desirable to use a material or multiple materials that
have magnetic phase transitions that span the required temperature range. In early
studies of regenerative magnetic cooling cycles, and in patents that resulted, lists
of ferromagnets with Curie points that vary between liquid helium temperature and
room temperature may be found [40, 41]. One may alloy Gd with, for example, Y,
30 Magnetocaloric Materials and Applications 1501

to tune the Curie temperature below room temperature. The Curie temperature of
Gd (294 K) is suppressed by approximately 2 K per percent of Y substitution [42,
43]. Other Gd-R (where R is a rare earth) alloys with reduced Curie points have
been investigated in terms of their magnetocaloric performance and even trialled in
cooling engines (see [44] for a review). However, there are very few Gd alloys that
have ordering temperatures above that of elemental Gd, with the notable exceptions
being Gd6 Mn23 (ferromagnetic, with Curie temperature 461 K) [45], the Gd-Co
alloy system (ferromagnetic, with Curie temperatures up to 1000 K), and Gd5 Si4
(ferromagnetic, with Curie temperature 336 K) [46]. The last of these materials
was the source of renewed interest in magnetic cooling following the observation,
by Vitalij Pecharsky and Karl Gschneidner Jr., of a giant magnetocaloric effect
(GMCE) in Gd5 Si2 Ge2 in 1997. The key to their discovery of a GMCE was a phase
transition that was not continuous (second order) but instead was discontinuous (first
order), mirroring the physics of the liquid-gas phase transition.
The observation of magnetocaloric effects in the neighborhood of first-order
magnetic transitions is essential to the development of new materials that satisfy
more of the requirements for application outlined above. Principally, the need for a
tunable Curie point means that alloying is typically required. However, any dilution
of the magnetic species can result in a reduction of the magnetocaloric effect, since
all atoms in the refrigerant need to be heated and cooled, not just those that are
associated with magnetic (dis)ordering. This point is especially vital when trying to
reduce refrigerant cost and criticality by moving away from rare earth gadolinium
to 3d metal-based magnetic refrigerants. Since the ordering temperatures of Fe,
Ni, and Co are much higher than room temperature, significant dilution (alloying)
is required. In order to offset the increase in heat capacity per unit volume that
comes from alloying, an increase in ∂M ∂T beyond that associated with simply using
a continuous phase transition is required. For this reason, GMCEs have all been
associated with first-order magnetic phase transitions. The following subsections
summarize the magnetocaloric effects associated with some key examples of
materials with first-order magnetic transitions near room temperature. For recent
work on the large magnetocaloric effect associated with the second order transition
in AlFe2 B2 -based compounds, we refer the reader to [180,181,182].

The Gd5 (Si,Ge)4 System


Gd5 Si2 Ge2 , which reignited research in the field of room temperature magne-
tocaloric materials, is to be found in the middle of the Gd5 Si4 -Gd5 Ge4 phase
diagram (Fig. 5). As is typical of giant magnetocaloric materials, we immediately
see that there is a competition between multiple magnetic and structural phases.
This complexity is a source of enhanced magneto-elastic coupling and therefore
larger MCEs. In the case of Gd5 Si2 Ge2 , the low temperature state is ferromagnetic
and monoclinic, while the high temperature state is paramagnetic and orthorhombic.
This compound exists in a window of the (Si,Ge) solid solution where the magnetic
and structural transitions coincide. At the transition, one half of the Si-Ge bonds that
join slabs of material break [47], an example of a symmetry-changing, displacive
transition and is also associated with a large thermal hysteresis of about 10 K.
1502 K. G. Sandeman and S. Takei

350

300
O(II ) M O(I )

250

200
TN, TC(K)

150

100

, , , Tc
50 TN

Gd5(SixGe1-x)4
0
0.0 0.2 0.4 0.6 0.8 1.0
Gd5Ge4 x(Si) Gd5Si4

Fig. 5 Magnetic and structural phase diagram of the Gd5 Si4 -Gd5 Ge4 system. We note the
large range of competing magnetic states (separated by solid and dashed lines) and structural
states (separated by dot-dashed lines) and the region of greatest interest for room temperature
magnetocaloric studies, the Gd5 Si2 Ge2 composition, that exhibits a magneto-structural phase
transition involving magnetic ordering and a change of crystal symmetry. O(H) phase properties are
indicated by circles, while triangles are for the M phase and squares for the O(I) phase. (This figure
is reprinted from Journal of Alloys and Compounds, 260, V. Pecharsky and K.A. Gschneidner Jr.,
“Phase relationships and crystallography in the pseudobinary system Gd5 Si4 -Gd5 Ge4 ”, 98–106,
Copyright 1997, with permission from Elsevier [49])

For this reason, while the isothermal entropy change associated with field-induced
increases in magnetization is large, the adiabatic temperature change is not wholly
reversible [108]. Initially reported maximum values were around 7 K in 0–2 T and
around 15 K in 0–5 T [48]. At this point, it is worth noting that there is currently no
agreed definition of a “giant” magnetocaloric effect. In Gd5 Si2 Ge2 , an isothermal
entropy change of around 20 Jkg−1 K−1 was observed in a field change of 0–
5 T. This level of entropy change is normally termed “giant,” although it does not
guarantee that there is also a large reversible Tad (T).
Subsequent work on the Gd5 Si4 -Gd5 Ge4 system examined how to tune the
material so as to reduce the thermal hysteresis at the magnetic ordering transition
while maintaining the MCE. It was found that small amounts of impurities drove
30 Magnetocaloric Materials and Applications 1503

the system toward a second-order Curie transition, eliminating the thermomagnetic


hysteresis but reducing the MCE [50, 51]. As is common to many room temperature
magnetocaloric materials, the presence of multiple interactions and first-order
transition(s) has led to some valuable physical insights stemming from the coupled,
magneto-structural transition that involves a change of crystal symmetry and
volume. The large volume change allows for giant magnetostriction to be observed
in single crystals of the room temperature magnetocaloric Gd5 Si2 Ge2 phase [52];
it has arrested dynamics which give rise to the possibility of thermoelectric effects
such as spontaneous voltage generation [53, 54]; and in thin films, the incomplete
nature of the phase transition when the material is subject to substrate strain yields
both decreased volumetric S(T) values and also reduced transition hysteresis [55].

The (Mn,Fe)2 (P,Si) System


The set of Mn-based ferromagnets that are amenable to magnetocaloric investiga-
tions is very large. Some, such as vacancy-doped or chemically substituted MnCoGe
(sometimes written as CoMnGe), crystallize in either the orthorhombic Pnma space
group or in the hexagonal P63 /mmc space group. They can even exhibit magnetic
order-disorder transitions between the two structure types simultaneously, in a way
analogous to the Gd5 Si4 -Gd5 Ge4 system. Large isothermal S(T) values have been
found in such cases (see Refs. [56–60] and the subsection on martensitic materials
later). However, as with Gd5 Si4 -Gd5 Ge4 , the displacive transition poses a problem
for hysteresis minimization. Without simultaneous optimization of interfacial strain
(see later), or without a very large applied magnetic field of 7 T [60], thermal
hysteresis at the transition prevents the magnetocaloric effect from being completely
reversible [57, 61].
This hysteresis problem has been effectively dealt with by recourse to a magneto-
elastic rather than magneto-structural first-order phase transition, in (Mn,Fe)2 (P,Si),
a material which belongs to the Fe2 P family (space group P-62m). Fe2 P is a strongly
anisotropic hexagonal ferromagnet with a first-order ferromagnetic ordering tem-
perature of 216 K [64, 65]. This transition involves a change of the c/a ratio of the
hexagonal unit cell and has only a small thermal hysteresis associated with it. Fe2 P-
based alloys were originally studied in a room temperature magnetocaloric context
by Ekkes Brück’s group in the early 2000s.
The first magnetocaloric studies on Fe2 P-based compounds were on MnFe(P,As),
in which a GMCE comparable to that seen in Gd5 Si2 Ge2 could be found (see
Fig. 6a) [62]. However, the urge to reduce the arsenic content has driven a
large effort to control and optimize the variation of Curie temperature, magne-
tocaloric effect, and magnetothermal hysteresis by doping. The resulting alloy
series, (Mn,Fe)2 (P,Si), is a hexagonal ferromagnet that exhibits so-called mixed
magnetism; the Mn atoms sit on the 3f site of the Fe2 P structure, while the Fe
atoms sit on the 3g site. However, for optimal hysteresis, this occupation is not
necessarily equal nor full (e.g., Mnx Fe1.95−x P1−y Siy [63]; see Fig. 6b). The physical
origin of this optimization is not fully known. However, it has been shown using
density functional theory simulations [63, 66] that there is a delocalization of the
3g site moment upon entering the paramagnetic state of both the (Mn,Fe)2 (P,Si) and
1504 K. G. Sandeman and S. Takei

a b
20 20
MnxFeP1-95-xP1-ySiy
Gd5Ge2Si2 MnFeP0.45As0.55

15 15
- Δ Sm (Jkg-1 K-1)

- Δ Sm (Jkg-1 K-1)
10 10

5 Gd 5

0 0
260 280 300 320 340 200 250 300 350 400
T (K)
T (K)

Fig. 6 (a) Isothermal entropy curves of MnFe(As,P), where a favorable comparison with
Gd5 Si2 Ge2 and with Gd can be made. Reprinted by permission from Springer Nature: Nature,
415, 150, O. Tegus et al., copyright 2002 [62]. (b) The arsenic in this material series has been
successfully replaced by Si to provide large magnetocaloric effects in Mnx Fe1.95−x P1−y Siy . Again
the S(T) for Gd is shown for comparison. (This figure is copyright © 2011 WILEY-VCH Verlag
GmbH & Co. KGaA, Weinheim and is reproduced from [63] by permission of Wiley. In the plot,
0–1 T (open symbols) and 0–2 T (solid symbols) effects are shown for x = 1.34, 1.32, 1.30, 1.28,
1.24, 0.66, 0.66 and y = 0.46, 0.48, 0.50, 0.52, 0.54, 0.34, 0.37 (left to right))

the parent Fe2 P compound, causing a loss of magnetic moment. This results from
covalent bond formation between Fe and P/Si. By comparison, the 3f site moment is
relatively well maintained into the paramagnetic state. Consequently, the action of a
magnetic field not only orders the magnetic moments but increases their size. Such
mixed magnetism physics is not unique to (Mn,Fe)2 (P,Si), but it clearly improves
the MCE.

The La(Fe,Co,Mn,Si)13 Hy System


As in the case of (Mn,Fe)2 (P,Si), the compounds in the La(Fe,Co,Mn,Si)13 Hy
material family that are relevant to room temperature cooling are based on a material
that has a lower Curie point: La(Fe,Si)13 . This compound, which has the NaZn13
structure, is formed in a peritectic reaction and is rarely available in single-crystal
form. Microstructural studies have also sought to address, and minimize, the amount
of alpha-Fe phase present in the final material through optimization of annealing
temperature and Si content [67]. The La(Fex Si1-x )13 phase possesses a giant MCE
which has a first-order, iso-symmetry magnetic ordering transition for x > 0.86
that decreases in temperature with the replacement of silicon by iron, e.g., from
around 210 K for x = 0.86 to around 195 K for x = 0.88, corresponding to
a ∂TC /∂x of around 7.5 K per percent of Fe/Si substitution [68, 69]. The first-
order transition is furthermore associated with a change of volume that depends on
composition and proximity to the tricritical point. (Compounds in the La(Fex Si1-x )13
family with 0.82 < x < 0.86 have second-order Curie transition temperatures above
210 K [69].)
30 Magnetocaloric Materials and Applications 1505

The ferromagnetic ordering temperature can also be increased to room tempera-


ture and beyond by Fe/Co substitution or by interstitial hydrogenation. The action of
Fe/Co substitution is to drive the phase transition second order, whereas full hydro-
genation, which is empirically found to be around y = 1.6 for La(Fe0.88 Si0.12 )13 Hy ,
increases the unit cell volume by 3% and is the only interstitial doping that preserves
the first-order nature of the ordering transition while raising the Curie temperature.
Indeed, as with (Mn,Fe)2 (P,Si), the magnetic ordering temperature can be raised
well above room temperature, to 336 K in the case of La(Fe0.88 Si0.12 )13 H1.6 at
full hydrogenation [70] and even higher for compounds with lower values of iron
content x. The strong dependence of ordering temperature on hydrogenation is
understood to be a result of the action of the lattice expansion on the density of elec-
tronic states. The La(Fex Si1-x )13 system of ferromagnets exhibits itinerant electron
metamagnetism (IEM) and strong spin fluctuations near the Curie point [71, 72].
The action of hydrogenation is understood to only mildly modify the electronic band
structure, compared to the greater action of other interstitials such as C or N [73].
There has been a wealth of fundamental and applied studies on the La(Fe,Co,
Mn,Si)13 Hy system. Early studies of partially hydrogenated La(Fe0.88 Si0.12 )13 H1.0
showed excellent performance in a prototype cooling engine, beating the traditional
refrigerant Gd and even Gd0.94 Er0.06 (which has a lowered Curie point) at tem-
peratures close to 283 K [74]. However, partially hydrogenated materials are not
satisfactory because it is found that, if stored at their phase transition temperature,
they undergo magnetic phase separation, due to internal migration of hydrogen [75].
While the hydrogen can be rehomogenized by aging for a few minutes away from
the phase transition, the magnetic phase separation is a source of irreversibility that
makes application of the material unattractive. The dynamics of hydrogen pumping
between phases when the sample is in a mixed state has been studied, showing
that partially hydrogenated material can phase segregate on a timescale of hours
[76]. For this reason, full hydrogenation is currently preferred, in combination with
Fe/Mn doping (which lowers the ordering temperature as required to the desired
value) [77, 78]. The resulting material is stable over time as the hydrogen is much
less mobile.

Fe-Rh
While all of the GMCEs presented so far in this section have been associated with
ferromagnetic ordering transitions, there is an example of a first-order, magnetic
order-order transition that is also associated with a giant MCE. The Fe-Rh binary
phase diagram contains a small region near to the equiatomic FeRh composition
which crystallizes in a cubic CsCl structure and has an antiferromagnetic (AFM)
ground state. Upon raising temperature, this ground state gives way to a ferro-
magnetic (FM) state, and the resulting metamagnetic transition between the two
states is first order. This phase transition has been the subject of applied research
interest in the field of heat-assisted magnetic recording and the focus of fundamental
interest for more than 50 years [183]. As with the case of (Mn,Fe)2 (P,Si), there is
the development of an increased moment in one phase: in this case, the Rh atom
takes on a ∼ 1 μB moment in the ferromagnetic state. This, combined with the large
1506 K. G. Sandeman and S. Takei

saturation magnetization and low dilution of magnetic species, yields a material that
has the largest Tad values (per Tesla applied field) of any MCE compound. In a
field of 2 T, the single-shot adiabatic temperature change reaches 12 K [79].
The reversibility of the MCE in Fe-Rh has been a subject of debate [80–84].
Manekar and Roy have presented a comprehensive discussion of the impact of the
thermomagnetic hysteresis at the AFM-FM transition on the reversibility of the
GMCE [80]. Instead a particular path in field and temperature space has to be chosen
for a giant MCE to be observed after a first adiabatic application of a magnetic
field. Whereas early studies were concerned with the diminution of the MCE with
magnetic field cycling [82, 83], recent studies have tried to ascertain the reversible
component and have surmised that it is 15% higher than in Gd [85]. Work by the
Mañosa group has confirmed that, within a given window of temperature, both the
magnetocaloric effect and the barocaloric effect (using an applied pressure to couple
to the volume change at the AFM-FM transition) are giant and reversible [84]. The
two effects can even be combined to yield large so-called multicaloric effects [81].
FeRh arguably highlights the potential for yet greater Tad to be found in d-metal
materials. A summary of the peak values of Tad and S is given in Fig. 7 using
reversible Tad unless noted. As Wood and Potter [86] found, a material coefficient
of performance can be constructed by examining the ratio of Tad × S to the input
magnetic work. They evaluated the resulting coefficient of refrigerant performance
(CRP) for a number of second-order materials. As many of the available materials
have somewhat similar saturation magnetizations, we chose to show Tad and
S as a guide to the state of material development and in order to demonstrate
avenues for exploration. In Fig. 7 we see that the two d-metal material systems that
are closest to commercialization, La(Fe,Co,Mn,Si)13 Hy and (Mn,Fe)2 (P,Si), exhibit
similar values of Tad (6 K in a 0 to 2 T field change). Fe-Rh has a considerably
higher peak Tad value, and it indeed the quantity μ0 −1 ∂Tt /∂H has a close to
optimal value in this case. (For a fixed saturation magnetization, there is a value
of ∂Tt /∂H or ∂TC /∂H that maximizes the adiabatic temperature change. However,
this value of ∂Tt /∂H or ∂TC /∂H will not simultaneously maximize the isothermal
entropy change. For more details, see refs. [87,88]). We note that, with the exception
of Fe-Rh, there is plenty of room for optimization of the adiabatic temperature
change.

Heusler Compounds and Other Martensitic Materials


The last large group of MCE materials that we will discuss is materials exhibiting
a martensitic (diffusionless) phase transition, of which the most often studied in a
magnetocaloric context are alloys based on three Heusler compounds : Ni2 MnGa,
Ni2 MnSn, and Ni2 MnIn. All three alloys have a martensitic transformation from a
low-temperature martensite phase (tetragonal, L10 ) to a high-temperature austenite
phase (cubic, L21 structure) [89]. The martensitic transformation temperature can
be readily tuned by composition, and the appearance of ferromagnetism is likewise
strongly dependent on structure type, through composition. Conventional MCEs are
30 Magnetocaloric Materials and Applications 1507

16
Fe-Rh model
0 to 2 Tesla
Fe-Rh
Mn-Fe-P model
Fe-Rh
14 La-Fe-Si model
La-Fe-Si
Fe-Rh
12 increasing
| (K)

Mn-Fe-P | dTc/dH |

10 La-Fe-Si dTc
<
dH
max

8 La(Fe0.88Si0.12)13Hy Mn-Fe-P
GdSi2Ge2
| Tad

MnFe(As,P)
6 Gd La(Fe,Mn,Si)H13
MnxFe1.95-x P1-ySiy
La(Fe1-xCox)11.9Si1.1
4

2 La1-xAgyMnO3
La0.6Ca0.4MnO3
MnAs
0
0 5 10 15 20 25 30 35 40
max -1 -1
| S | (JK kg )
Fig. 7 A comparison of the isothermal entropy and adiabatic temperature change found in a
selection of room temperature magnetocaloric materials, after Ref. [88]. Where available, the
relevant reversible effect is shown in 2 T. This has a deleterious effect on the magnitude of the
effect seen in MnAs, in particular. (The exception is FeRh, where the irreversible effect is shown
to draw attention to the possibility of higher Tad values in future compounds.) The solid line gives
a theoretical limit to the combination of (S, Tad ) available in a typical transition metal alloy of
saturation magnetization 120 Am2 /kg. (This figure is reprinted from Scripta Materialia, 67, K.G.
Sandeman, “Magnetocaloric materials: The search for new systems”, 566–571, Copyright 2012,
with permission from Elsevier)

associated with the Curie point in ferromagnetic Ni2 MnGa, an alloy best known for
the shape memory effect associated with its martensitic transformation. It is possible
to merge the martensitic and Curie transformations [90], but doing so generally
brings about a strongly first-order phase transition with large thermal hysteresis,
and an inability to drive the transition out of any remanent metastable state solely
with a magnetic field since the field sensitivity of the martensitic transition is
low.
Heusler compounds have been a fertile playground in which to explore inverse
magnetocaloric effects, with giant inverse effects first being reported by Krenke
et al. in Ni-Mn-Sn at room temperature [91]. These effects were associated with re-
entrant ferromagnetism – the Curie point of the low-temperature martensite; TC M
is lower than that of the austensite (TC A ), and so, on heating through the structural
transition, ferromagnetism emerged. Such an inverse MCE is common to both Ni-
Mn-Sn and Ni-Mn-In alloys, as can be seen from the relative positions of TC M and
TC A in the magnetic phase diagrams shown in Fig. 8.
1508 K. G. Sandeman and S. Takei

x (at%)
20 10 0 20 10 0
1200
Ni50Mn50-xSnx Ni50Mn50-xInx Ni-Mn-Ga
1000

MS MS MS
800 Cubic Cubic Cubic
(K)

PM PM PM
600
T

L10 L10 A L10


A A TC
400 TC TC

14M
200 T CM T1
T CM
FM (a) FM (b) (c)
0
7.5 8.0 8.5 7.5 8.0 8.5 7.5 8.0 8.5
10M e/a 10M 14M
14M
10M

Fig. 8 Magnetic and structural phase diagrams for the three Heusler alloys most studied in the
context of the magnetocaloric effect: Ni-Mn-Sn (a), Ni-Mn-In (b), and Ni-Mn-Ga (c). Magnetic
and structural (martensitic) transformations are indicated by circles and triangles, respectively.
Dotted lines demarcate regions of each phase diagram with different (modulated) structures.
Premartensitic transition temperatures are shown as small solid circles in (c). (Figure reproduced
from [89] with permission from IOP Publishing)

In addition to studies of conventional and inverse MCEs, the Heusler compounds


have provided the opportunity to test theories of thermal hysteresis control; ab
initio models of magnetism at finite temperature; and proposals to use multicaloric
approaches to circumvent problems associated with thermal hysteresis in strongly
first-order transitions. A prediction that thermal hysteresis at a martensitic transfor-
mation is minimized by optimization of the austenite/martensite interface through
so-called co-factor conditions on the eigenvalues of the transformation matrix
[92] has been realized in a series of shape memory materials and non-magnetic
martensites [93, 94]. Quantum mechanical ab initio models have successfully
reproduced the phase diagram of Ni-Mn-Ga and, in particular, trends in the
martensitic temperature with chemical doping [95]. Most recently, a stress and
magnetic field have been combined to provide a proof-of-concept multicaloric cycle
that provides a “hysteresis-positive” approach to the use of strongly first-order
materials; the retention of a metastable magnetic state upon removal of a stimulus
is used in designing the final cycle [33].
Finally, it is worth noting that martensitic phase transformations are common
in Mn-based intermetallics. Large, albeit irreversible, MCEs have been associated
with the orthorhombic-hexagonal transition in CoMnGe upon compositional tuning
(e.g., in [96]) and over a wide range of temperatures in Ni(Mn,Fe)Ge [97].
30 Magnetocaloric Materials and Applications 1509

Room Temperature Magnetic Cooling Engines

Of the materials outlined above, the (Mn,Fe)2 (P,Si) and the La-Fe-Si series are
closest to commercialization at the time of writing. While this chapter does not
examine the nature of the cooling engine itself, it is worth noting that almost
all magnetic cooling engine prototypes documented in a review by Yu et al. in
2010 used gadolinium or gadolinium alloys [98]. The relative availability of Gd
and its large MCE make Gd a good material with which to test engineering
models, thermodynamic cycles, and regenerator conformations. In all prototypes,
the refrigerant and magnet must undergo relative motion, and a fluid heat exchange
medium is circulated over the solid refrigerant. Early prototype ideas used either
linear actuation, such as Brown’s prototype [34] or rotational motion, such as was
employed in Steyert’s design [38] and in the prototype of Kirol and Dacus [178,
179].

Issues for Deployment of Known Magnetocaloric Materials in Cooling


Engines
Firstly, it is perhaps straightforward to note that the final morphology of the working
material in a magnetic cooling engine has to be chosen to allow for fast heat
exchange with whatever fluid is used to transport heat. This means that thin plates
or fine, porous structures of magnetocaloric material need to be produced. As we
have noted, all of the GMCE materials possess a magneto-elastic or magneto-
structural phase transition that involves a change of crystal lattice parameters or
even crystal structure. The inter-grain strain polycrystalline materials – the most
common form of magnetocaloric sample – can potentially result in cracking upon
multiple field cycles, something which has been shown to be moderated by, for
example, the substitution of boron for phosphorus in the (Mn,Fe)2 (P,Si) class [99]
or by strain relief of magnetocaloric particles in a porous matrix [100]. A solution
that maximizes the magnetocaloric effect per unit volume is preferable, since it is
the volume of magnetic field energy required (typically provided by a permanent
magnet for small power applications) that is the primary cost factor in the final
device.
One final point of interest in this brief overview of GMCE materials is the
variation of transition temperature with composition. In (Mn,Fe)2 (P,Si) materials,
∂TC /∂xSi is approximately 10 K per percent of Si replaced, while the thermal
hysteresis associated with the phase transition varies by at least 4 K per percent
of Si [63]. In La-Fe-Si, a rate of ∂TC /∂xSi ∼ 5 K per percent of Si is found in
the hydrogenated series. Deployment of either material will require close process
control of composition, minimizing or at least accounting for any changes that occur
during the production of material parts.
Before moving on to some alternative uses and types of caloric effects, we
present a summary table of magnetocaloric and giant magnetocaloric effects in
Table 1. A similar table is to be found in [101]; here we have added some further
compounds from the families presented above. Where possible, reversible effects in
1510

Table 1 Example compounds from various families of magnetocaloric material, with MCEs from low temperature to room temperature
Peak |S| in
Material family Type of transition Transition temperature (K) 0–1 T(Jkg−1 K−1 ) Peak |Tad | in 0–1 T (K) Refs.
Cu N/A (nuclear Microkelvin regime – – See text
paramagnet)
GGG N/A (para-magnet) Millikelvin regime – – See text
Gd 2nd order Curie 294 3.0 2.5 [86]
Gd5 Si2 Ge2 1st order Curie 272 11 3.5 (irrev.) [47]
[Gd5 (Si,Ge)4 family]
LaFe11.384 Si1.26 Mn0.356 H1.52 1st order Curie 290 11 3.5[in 1.2 T] [102]
[La(Fe,Si)13 family] [in 1.2 T]
MnFe0.95 P0.605 B0.065 Si0.33 1st order Curie 274 (warming) 12.5 2.4 [103]
[(Mn,Fe)2 (P,Si) family]
MnAs 1st order Curie 318 (warming) 32 (irrev.) 2.5 (irrev.)0.2 (rev.) [104][105]
316 (warming)
Fe49 Rh51 1st order AFM-FM 310 (warming) 12 (rev.) –6.2 (rev. in 2 T) [106][85]
∼12 (rev.)
Ni2.19 Mn0.81 Ga 1st order Curie 338–348 (warming) 5.5 (rev.) 0.6 (after 6 cycles to 1.85 T) [107, 108]
[Ni-Mn-Ga family]
Ni45.7 Mn36.6 In13.5 Co4.2 1st order 287 (max. irrev. effect) 13.8 (full transition) 8 (irrev.)3 (rev.)[in 1.95 T] [109]
[Ni-Mn-In and Ni-Mn-Sn AFM/PM-FM 282 (max. rev. effect)
families]
Data shown are the transition temperature of relevance (if any); peak values of Tad and S in a field change of 0–1 T (where known); and references to relevant
measurements. Where possible, reversible values of Tad and S are shown; in the case of MnAs and the inverse, both irreversible (irrev) and reversible (rev)
values are given. For the NDR and ADR refrigerants, Cu and GGG, respectively, cooling is associated with no change of magnetic state, and therefore no data
associated with a magnetic transition are noted. We also note that all of the materials with first-order transitions can be driven second order by some change in
composition, external field; for this reason a sample of only their first-order properties are reported below since those have formed the focus of research in each
case
K. G. Sandeman and S. Takei
30 Magnetocaloric Materials and Applications 1511

a field change of zero to 1 T are given. As there is not yet a consistent approach
to the measurement of the MCE in the literature, where data is not available in the
same field change, this is noted. We note that such a table can only present a small
fraction of the total number of materials studied. For this reason, we have focused
on presenting material families as in the above subsections, along with appropriate
references for the reader to begin further research.

Power Conversion and Spin Caloritronics

We finish with some remarks about two developing areas that relate to the reverse
of the magnetocaloric effect. Instead of generating a thermal gradient from an input
of magnetic work, both of these areas use a thermal gradient to generate either a
change in static magnetization via the pyromagnetic effect or a spin current via the
spin Seebeck effect.

Power Conversion via Pyromagnetic Materials

The first area is the pyromagnetic effect (PME), whereby input heat is converted
into work, rather than vice versa. This “magnetocaloric effect in reverse” arguably
predates the magnetocaloric effect as it was considered by both Tesla and Edison in
the late nineteenth century as a means of generating power from heat.
Applied studies of pyromagnetic energy conversion can be divided into two main
types. The first is direct conversion of a thermal gradient to electrical energy, for
example, through the change of magnetic flux,  through a coil that has in its core
a working (pyromagnetic) material that undergoes a change of magnetization when
heated. Tesla first proposed this in 1889 [110]. The pyromagnetic voltage, V, is then

δ
V =− ; = B.dA
δt coil

The second type is conversion of a thermal gradient to mechanical energy, for


example, by the movement of the pyromagnetic material via decreased attraction to,
or repulsion from, a permanent magnet as the former is heated. The first of Thomas
Edison’s two patents in the area was of this type [111] and suggested heating iron
through its Curie temperature. (His second patent employed induction, combined
with a thin tube morphology of the iron to improve heat exchange [112].) One of
the most common designs for conversion of heat into mechanical energy is the Curie
wheel, a concept that was patented first by Bremer [113]. It involves the use of a ring
of working pyromagnetic material that is rotated through a field source or series of
field sources, where heat is applied to the working material at points that are close
to the areas of high field. At the time of writing, the only kilowatt prototype known
to the authors is that of a Swiss company using a Curie wheel design, employing a
ferromagnet that is heated through its Curie point [114].
1512 K. G. Sandeman and S. Takei

Fig. 9 A schematic power generation cycle in M(H, T) space based on the conversion of heat
energy using a pyromagnetic working material that changes magnetization with temperature. In the
case shown, the working material is ferromagnetic at the cold sink temperature and paramagnetic
at the hot sink temperature. (Figure reproduced from C.-J. Hsu, S. M. Sandoval, K. P. Wetzlar,
and G. P. Carman, J. Appl. Phys. 110, 123923 (2011) [115], with the permission of the American
Institute of Physics)

No matter whether electrical or mechanical energy is produced, the measure


of efficiency of a thermomagnetic heat engine is the ratio of work output to heat
input. The work output that can be harnessed from a particular material can be
estimated from the area between the magnetization isotherms, M(H, T), at the hot
and cold temperatures of the thermodynamic cycle, calculated in a clockwise loop
(see Fig. 9). Furthermore, the maximum efficiency of a cycle based on a temperature
difference (e.g., between a waste heat source at Thot and a heat sink at Tcold ) is the
Carnot efficency, ηCarnot .
From basic thermodynamic principles, the efficiency of any heat to power
conversion increases as the available temperature difference (between the hot and
cold sinks at temperatures Thot and Tcold , respectively) increases. Indeed, the Carnot
efficiency is given by

Tcold
ηCarnot = 1 − .
Thot

Just as in the case of room temperature magnetic refrigerants, it has been


suggested as early as in Edison and Tesla’s early patents that the optimization of
pyromagnetic effects may be brought about by use of a phase transition. In this
instance, the temperature of any transition should be below the temperature of the
heat source. Both of the current candidate room temperature magnetic refrigerant
materials can have Curie temperatures well above room temperature: by fully
hydrogenating La-Fe-Si or by varying the P/Si ratio in (Mn,Fe)2 (P,Si) as described
earlier. Hence, they may be considered for the purpose of power conversion from
30 Magnetocaloric Materials and Applications 1513

waste heat. However, these materials are by no means the only options for use as
power conversion materials [184]; for example, it has been shown that a martensitic
Ni-Mn-Ga thin-film cantilever can be flexed to produce milliwatts of power [116].
While in a cooling cycle a latent heat may be a desirable property of a working
material (where heat transfer is an output), a latent heat actually provides an
additional heating requirement in the case of a heat engine (where heat transfer is an
input). For this reason, there is ongoing discussion as to the optimal requirements
for a pyromagnet in a heat engine, in particular due to the trade-off between power
output and efficiency that occurs when the temperature span of the pyromagnet is
varied [117].
We also note that, while the Carnot efficiency of a power conversion cycle
is optimized by reducing the cold sink temperature, it has been shown that the
percentage of Carnot efficiency achievable with a pyromagnet increases as the
temperature difference between Thot and Tcold decreases [115]. There is therefore
a balance to be struck between the Carnot efficiency and the achievable efficiency.

Spin Seebeck and Spin Nernst Effects and Spin Caloritronics

Rapid progress in spintronics was triggered by the discovery of spin Hall phe-
nomena in heterostructures consisting of interfaces between a magnetically ordered
material and a non-magnetic metal with strong spin-orbit coupling (SOC). SOC
(in the metal and at the interfaces) together with exchange coupling (between the
conduction electron spins and the magnetization) at a normal metal/magnet interface
allows angular momentum to be transferred from the metal’s crystal lattice to the
magnetization and engenders macroscopic coupling between the electrical current
in the metal and the magnetic order parameter. These magnetic heterostructures
have played a central role in modern spin caloritronics, where they are subjected
to an additional thermal gradient to study the coupled propagation of charge,
spin, and heat through them [118–120]. A spectacular development in the field of
spin caloritronics came from the discovery of the spin Seebeck effect (SSE) [121]
in which spin currents are thermally excited by a temperature gradient across a
magnetic system. The spin currents are then injected into an adjacent spin-orbit
coupled metal, converted into a charge current or an electromotive force transverse
to the current direction via the inverse spin Hall effect (SHE) and electrically
detected. From the viewpoint of device applications, the field has strong relevance,
as it may lay the foundations for new devices with superior thermoelectric figures of
merit and more efficient energy harvesting as well as new means to manage immense
Joule heating expected in future high-density silicon electronics [122].
The SSE has been studied using two configurations depending on the relative
orientation of the spin current with respect to the thermal gradient: longitudinal
[123–132] (see Fig. 10a) and transverse [123, 133–135] (see Fig. 10b). Historically
speaking, the transverse configuration was investigated first, but the current status
of the field is such that the physics of the longitudinal SSE is more well-understood
relative to its transverse counterpart. Thus far, experiments have focused on insulat-
1514 K. G. Sandeman and S. Takei

Fig. 10 Two configurations with which spin Seebeck effect has been investigated: (a) the
longitudinal and (b) the transverse. Spin current is denoted above by Js , and the transverse
electromotive force generated by the spin current via the inverse spin Hall effect is labeled
by EISHE . (Figure reproduced from K. Uchida et al., J. Appl. Phys. 111, 103903 (2012) with
permission of the American Institute of Physics)

ing magnets, as longitudinal SSE signals are indistinguishable from classical Nernst
signals if ferromagnetic conductors are used [136, 137]. We will therefore begin by
focusing on insulating ferromagnets and then comment on antiferromagnets later.
General theories of the longitudinal SSE have been discussed in refs. [138–142].
The phenomenon can be explained in terms of a spin current inside the ferromagnet,
generated by magnetic quasiparticles (i.e., magnons) drifting under the thermal
gradient, that is, injected into the metal and converted into a transverse electric field
EISHE via the inverse SHE,

EISHE = θSH ρJs × σ. (10)

Here, θ SH is the spin Hall angle quantifying the efficiency of conversion between
spin and charge currents, ρ is the electrical resistivity of the metal, and Js is the
spatial direction of the spin current polarized along σ . Time-averaged spin current
Jsp pumped into the metal from the magnet by magnetic thermal noise is given by
spin pumping and reads [143]:


Jsp σ = Gr m × ṁ + Gi ṁ , (11)
2e

where  . . .  denotes time averaging, m is the unit vector parallel to the mag-
netization, and Gr or Gi (in units of −1 ) is the real or imaginary part of the
spin mixing conductance [144] that determines the magnitude of the pumped spin
current. Thermal spin fluctuations in the metal also subject the magnetization to a
stochastic spin torque, which can be expressed using a stochastic magnetic field h(t),
and give rise to time-averaged spin current Jn polarized normal to the magnetization
30 Magnetocaloric Materials and Applications 1515

Jn σ = − m × h , (12)


with the Cartesian components of h(t) obeying hi (t)hj (t ) = (Gr /2π )kB Tn δ ij

δ(t − t ) (Tn being the metallic electron temperature) in accordance with the
fluctuation-dissipation theorem. When the ferromagnet and the metal are in thermal
equilibrium, there is no net flow of spin across the interface, and we have
Js = Jsp + Jn = 0 [126]. However, since Jsp and Jn are proportional to the magnon
temperature Tm and Tn , respectively, a difference in the two temperatures destroys
the balance, and a net spin current Js ∝ Tm − Tn results [138–141]. In particular, one
notes that a sign reversal in the inverse SHE signal occurs as a result of reversing
the temperature gradient across the device. This was indeed observed in one of the
first longitudinal SSE experiments [130].
The temperature difference Tm − Tn at the detector end can be related to the
magnon spin current flowing toward the detector due to the thermal gradient [138].
One would expect that only magnons from distances smaller than their propagation
length contribute to the magnon spin current at the cold end and that this propagation
length depends on the quality factor of the ferromagnet (as spin damping would
lead to the decay of magnons into, e.g., the lattice). Reference [128] experimentally
determined this magnon propagation length ξ by studying the dependence of the
inverse SHE signal on the thickness of the magnetic film. The experiment was
conducted on a yttrium iron garnet (YIG)/platinum (Pt) bilayer. They observed
a saturation in the SSE signal once the thickness starts to overcome the mean
propagation length of the magnons and fitted their signal to

EISHE ∝ Tm − TN ∝ 1 − e−L/ξ . (13)

An upper limit for the propagation length of ξ ∼ 1 μm at room temperature


was in agreement with the effective magnon diffusion lengths extracted from time-
resolved measurements on the same bilayer material [129, 145]. By fitting Eq. (4) to
the detected inverse SHE signal over a temperature range, a monotonic increase in
ξ with decreasing temperature was also observed, ξ ∼ T−n with n ≈ 1 [146], which
agrees well with theoretical expectations [131, 147].
The magnetic field dependence of the SSE signal has also been investigated [127,
132]. The application of a magnetic field parallel to the equilibrium magnetization
leads to an energy gap in the magnon spectrum; thus an increase in the field should
suppress the thermal population of magnons and the SSE signal. Such a suppression
was indeed observed in the longitudinal SSE signal [127]; however, appreciable
magnon freeze-out was observed even at temperature scales much greater than the
energy gap [132]. As a possible explanation of this observation, [132] suggested
that magnons with energies below T ∼ 40 K contribute disproportionately to the
longitudinal SSE.
A new addition to the study of longitudinal SSE is antiferromagnets [148]. While
they cannot, in principle, harness the SSE due to complete magnetic compensation,
longitudinal SSE signals were reported for several antiferromagnets above the spin-
1516 K. G. Sandeman and S. Takei

flop transition [149, 150]. At low temperatures (T = 15 K) and under high magnetic
fields (B = 14 T), the spin Seebeck coefficient for MnF2 [150], in particular, was
calculated to be 41.2 μV/K, which is larger than even the largest values of the SSE
in bulk single crystal YIG at low temperatures. The antiferromagnet insulator NiO
has also emerged recently as an attractive system for longitudinal SSE studies [151,
152]. Non-trivial magnetic anisotropy in NiO lifts the degeneracy between the two
magnon modes associated with the two-sublattice structure and, therefore, permits
spin transport at small magnetic fields [148]. Reference [151] proposed that the SSE
signals are generated mainly by thermally driven magnons from the low-energy part
of the spectrum, consistent with the high-field suppression of the SSE mentioned
above.
In relation to the longitudinal SSE, the transverse SSE has been much more
controversial. The precise mechanism behind the transverse SSE is still not well
understood, mainly because of its sensitive dependence on the details of the
magnon/phonon temperature profiles across the devices and the possible role played
by phonons in the substrate on which the active device is grown [134]. Initially, an
unexpectedly high transverse SSE signal was observed in permalloy [153], where
thermally induced spin voltages extended over distances many orders of magnitude
larger than the spin diffusion length. Subsequently, a giant transverse SSE signal was
detected in YIG over several millimeters and was attributed to long-ranged magnon-
mediated spin transport resulting from YIG’s exceptionally high quality factor
(and the correspondingly long magnon diffusion lengths) [138, 139]. However,
an experiment on a ferromagnetic semiconductor GaMnAs reported a similarly
long-ranged SSE signal even after cutting the magnetic coupling in the material
along the direction of the thermal gradient [154], casting doubt on the long-ranged
magnon-mediated SSE idea. The giant enhancement of the SSE signal for YIG was
subsequently interpreted as a phonon-drag effect: the temperature gradient drives a
stream of phonons, which then drags the magnons and causes their convection [155].
This phonon drag idea was also supported by an experiment that involved a detail
of characterization of the temperature dependence of the spin-Seebeck coefficient,
magnetization, and thermoelectric power of the ferromagnetic semiconductor GaM-
nAs (in which transverse SSE was measured), alongside the thermal conductivity
and specific heat of the GaAs substrate on which the ferromagnet was grown [156].
In further experiments a curious sign reversal in the transverse SSE voltage signal
at a certain heater-detector distance was observed and was reported to depend
both on the YIG/heater interface opacity and the YIG thickness [135, 157]. The
results were qualitatively explained by a complex interplay between the temperature
dependences of the interfacial transparency, i.e., angular momentum transfer across
the YIG/metal interface and the diffusive properties of the thermally excited non-
equilibrium magnons [157], providing evidence for the magnon-mediated transverse
SSE.
The SHEs, collectively referring to a transverse pure spin current Js ∝ θ SH Jc × σ
generated by a longitudinal charge current Jc and the inverse effect (used extensively
in the detection scheme for the SSE), have fueled extraordinary advances in the field
of spintronics [158]. Here, θ SH is referred to as the spin Hall angle quantifying
30 Magnetocaloric Materials and Applications 1517

the charge-to-spin conversion efficiency, and σ is the unit vector parallel to the
spin current polarization. The spin Nernst effect (SNE) [159–161] refers to the
generation of a transverse spin current Js in the presence of a longitudinal heat
current Jh , i.e., Js ∝ θ SN Jh × σ , where θ SN is the so-called spin Nernst angle.
Very recently, the effect has been detected directly in platinum films by selectively
changing the spin transport boundary conditions at the platinum film boundaries
and recording the ensuing spin Nernst-driven changes in the thermopower in the

Transport Effect Detection


First detected optically in a GaAs semiconductor
(a) Spin Hall Effect strip by measuring the electron spin polarization
Js using magneto-optical Kerr microscopy [170].
Polarizations were normal to the sample plane and
were found to have opposite signs at the two
opposite edges of the sample parallel to the applied
current, consistent with the predictions of the effect
[163,164,171,172]. These findings were
corroborated around the same time in a 2D Rashba
−J c spin-orbited coupled semiconductor using
Generation of transverse spin current density circularly polarized electroluminescence [173].
Js with longitudinal charge current density Jc
due to spin-orbit coupling
First measured by injecting a pure spin current
(b) Inverse Spin Hall Effect from a permalloy film into platinum layer via spin
−J c pumping and observing the electrical voltaage
generated by the conversion of the spin current into
a transverse charge current [174]. An independent
demonstration of the effect was confirmed in a
diffusive conductor in a similar fashion by using a
ferromagnetic electrode to inject a spin-polarized
current [175].
Js
Generation of transverse charge current
density Jc with longitudinal spin current
density Js due to spin-orbit coupling
First detected directly in a platinum film by
(c) Spin Nernst Effect selectively changing the spin transport boundary
Js conditions at the platinum film boundaries and
recording the ensuing spin Nernst driven changes
in the thermopower in the direction along the heat
gradient [161]. Demonstration of the effect was
also reported in tungsten by observing the
magnetoresistance of the tungsten film in the
presence of a thermal gradient [162].
Jh
Generation of transverse spin current density
Js with longitudinal thermal current density Jh

Fig. 11 Various magneto-electric and thermal effects involving charge and spin and their first
experimental measurements. (Left column reprinted by permission from Springer Nature: Nature
Materials, 16, 977, S. Meyer et al., copyright 2017)
1518 K. G. Sandeman and S. Takei

longitudinal direction, i.e., along the heat gradient [162]. The effect may have also
been observed in tungsten as well [163].
New and interesting theoretical directions for realizing the SNE have been
proposed. In the intrinsic SHE, time reversal symmetry prohibits transverse charge
current but allows transverse spin current originating from Berry curvature of
electron bands [164, 165]. By analogy, it has been predicted that a temperature
gradient can induce a transverse magnon spin current in systems with magnon Berry
curvatures both in ferromagnets [166, 167] and antiferromagnets [168, 169] with
Dzyaloshinskii-Moriya interactions.
We conclude with a table (Fig. 11) that summarizes the magneto-electric
and thermal effects described in this section, along with their first experimental
measurements.

References
1. Wikus, P., Burghart, G., Figueroa-Feliciano, E.: Optimum operating regimes of common
paramagnetic refrigerants. Cryogenics (Guildf). 51, 555–558 (2011)
2. Pecharsky, V., Gschneidner, K., Pecharsky, A., Tishin, A.: Thermodynamics of the magne-
tocaloric effect. Phys. Rev. B. 64, 144406 (2001)
3. Morrison, K., Moore, J.D.: In: Gutfleisch, O., Sandeman, K.G. (eds.) Magnetic Cooling: From
Fundamentals to High Efficiency Refrigeration, 1st edn. John Wiley and Sons, Chichester (in
press)
4. Nikitin, S.A., Skokov, K.P., Koshkid’ko, Y.S., Pastushenkov, Y.G., Ivanova, T.I.: Giant
rotating magnetocaloric effect in the region of spin-reorientation transition in the NdCo5
single crystal. Phys. Rev. Lett. 105, 137205 (2010)
5. Takeuchi, I., Sandeman, K.G.: Solid-state cooling with caloric materials. Phys. Today. 68,
48–54 (2015)
6. Velders, G.J.M., Fahey, D.W., Daniel, J.S., McFarland, M., Andersen, S.O.: The large
contribution of projected HFC emissions to future climate forcing. Proc. Natl. Acad. Sci.
106, 10949–10954 (2009)
7. McLinden, M.O., Brown, J.S., Brignoli, R., Kazakov, A.F., Domanski, P.A.: Limited options
for low-global-warming-potential refrigerants. Nat. Commun. 8, 14476 (2017)
8. Goetzler, W., Shandross, R., Young, J., Petritchenko, O., Ringo, D., McClive, S.: Energy
Savings Potential and Opportunities for Commercial Building HVAC Systems. The report
was published by the Department of Energy in the US (2017)
9. Yu, B., Gao, Q., Zhang, B., Meng, X.Z., Chen, Z.: Review on research of room temperature
magnetic refrigeration. Int. J. Refrig. 26, 622–636 (2003)
10. Kitanovski, A., Tušek, J., Tomc, U., Plaznik, U., Ožbolt, M., Poredoš, A.: Magnetocaloric
Energy Conversion: From Theory to Applications, pp. 269–330. Springer, Cham (2015)
11. Smith, A.: Who discovered the magnetocaloric effect? Warburg, Weiss, and the connection
between magnetism and heat. Eur. Phys. J. H. 38, 507–517 (2013)
12. Thomson, W.: in Cyclopedia of the Physical Sciences, 2nd edn., J.P. Nichol (ed.). Richard
Green and Company, London and Glasgow, 838 (1860)
13. Weiss, P., Piccard, A.: Le phènomène magnètocalorique. J. Phys. Theor. Appl. 7, 103–109
(1917)
14. Debye, P.: Einige Bemerkungen zur Magnetisierung bei tiefer Temperatur. Ann. Phys. 386,
1154–1160 (1926)
15. Giauque, W.F.: A thermodynamic treatment of certain magnetic effects. A proposed method
of producing temperatures considerably below 1◦ absolute. J. Am. Chem. Soc. 49, 1864–1870
(1927)
30 Magnetocaloric Materials and Applications 1519

16. Giauque, W.F., MacDougall, D.P.: Attainment of temperatures below 1◦ absolute by demag-
netization of Gd2 (SO4 )3 .8H2 O. Phys. Rev. 43, 768–768 (1933)
17. De Haas, W.J., Wiersma, E.C., Kramers, H.A.: Experiments on adiabatic cooling of param-
agnetic salts in magnetic fields. Physica. 1, 1–13 (1934)
18. Tiselius, A.: Nobel Prize in Chemistry 1949 – Presentation Speech, Nobel Prize Chem. 1949 –
Present. Speech (1949)
19. Kurti, N.: The prehistory and early history of nuclear cooling: reflections of a cryodinosaur.
Czechoslov. J. Phys. 46, 3371–3375 (1996)
20. Spichkin, Y.I., Zvezdin, A.K., Gubin, S.P., Mischenko, A.S., Tishin, A.M.: Magnetic molec-
ular clusters as promising materials for refrigeration in low-temperature regions. J. Phys. D.
Appl. Phys. 34, 1162–1166 (2001)
21. Evangelisti, M., Brechin, E.K., Affronte, M., Okubo, M., Train, C., Verdaguer, M., Brechin,
E.K., Teat, S., Helliwell, M., Raftery, J., Evangelisti, M., Affronte, M., Collison, D., Brechin,
E.K., McInnes, E.J.L.: Recipes for enhanced molecular cooling. Dalt. Trans. 39, 4672
(2010)
22. Peschka, W.: Liquid Hydrogen: Fuel of the Future, pp. 17–70. Springer Vienna, Vienna (1992)
23. Pickett, G.R.: Cooling metals to the microkelvin regime, then and now. Phys. B. 280, 467–473
(2000)
24. Oja, A.S., Lounasmaa, O.V.: Nuclear magnetic ordering in simple metals at positive and
negative nanokelvin temperatures. Rev. Mod. Phys. 69, 1–136 (1997)
25. Bleaney, B., Lounasmaa, O.V.: Nuclear orientation and nuclear cooling experiments in Oxford
and Helsinki Part 2. Progress from 1945 to 1970. Notes Rec. R. Soc. 57, 323–330 (2003)
26. Gorter, C.: No title. Phys. Z. 35, 923 (1935)
27. Kurti, N., Simon, F.: Experiments at very low temperatures obtained by the magnetic method.
I. the production of the low temperatures. Proc. R. Soc. A Math. Phys. Eng. Sci. 149, 152–176
(1935)
28. Kurti, N., Robinson, F.N.H., Simon, F., Spohr, D.A.: Nuclear cooling. Nature. 178, 450–453
(1956)
29. Pobell, F.: Refrigeration by adiabatic nuclear demagnetization. Matter Methods Low Temp.,
Springer, Berlin. 181–225 (1996)
30. Knuuttila, T.A., Tuoriniemi, J.T., Lefmann, K., Juntunen, K.I., Rasmussen, F.B., Nummila,
K.K.: Polarized nuclei in normal and superconducting rhodium. J. Low Temp. Phys. 123,
65–102 (2001)
31. McMichael, R.D., Shull, R.D., Swartzendruber, L.J., Bennett, L.H., Watson, R.E.: Magne-
tocaloric effect in superparamagnets. J. Magn. Magn. Mater. 111, 29–33 (1992)
32. Tishin, A.M., Spichkin, Y.I.: The Magnetocaloric Effect and Its Applications. IOP Publishing,
Bristol (2003)
33. Gottschall, T., Gràcia-Condal, A., Fries, M., Taubel, A., Pfeuffer, L., Mañosa, L., Planes, A.,
Skokov, K.P., Gutfleisch, O.: A multicaloric cooling cycle that exploits thermal hysteresis.
Nat. Mater. 17, 929–934 (2018)
34. Brown, G.V.: Magnetic heat pumping near room temperature. J. Appl. Phys. 47, 3673 (1976)
35. Bahl, C.R.H., Nielsen, K.K.: The effect of demagnetization on the magnetocaloric properties
of gadolinium. J. Appl. Phys. 105, 013916 (2009)
36. van Geuns, J.R.: A study of a new magnetic refrigerating cycle. Philips Res Suppl. 6, 1 (1966)
37. van Geuns, J.R.: US Patent 3,413,814 (1968)
38. Steyert, W.A.: Stirling-cycle rotating magnetic refrigerators and heat engines for use near
room temperature. J. Appl. Phys. 49, 1216 (1978)
39. Barclay, J.A., Steyert, W.A.: 4332135 (1982)
40. Oesterreicher, H., Parker, F.T.: Magnetic cooling near Curie temperatures above 300 K. J.
Appl. Phys. 55, 4334–4338 (1984)
41. Reid, C.E., Barclay, J.A., Hall, J.L., Sarangi, S.: Selection of magnetic materials for an active
magnetic regenerative refrigerator. J. Alloys Compd. 207–208, 366–371 (1994)
42. Kito, S., Nakagome, H., Kobayashi, T., Saito, A.T., Tsuji, H.: In: Miller, S.D., Ross Jr., R.G.
(eds.) Cryocoolers, pp. 549–553. ICC Press, Boulder (2007)
1520 K. G. Sandeman and S. Takei

43. Lara Pérez, E.S., Betancourt, I., Hernández Paz, J.F., Matutes Aquino, J.A., Elizalde Galindo,
J.T.: Magnetocaloric effect in as-cast Gd1−x Yx alloys with x = 0.0, 0.1, 0.2, 0.3, 0.4. J. Appl.
Phys. 115, 17A910 (2014)
44. Balli, M., Jandl, S., Fournier, P., Kedous-Lebouc, A.: Advanced materials for mag-
netic cooling: fundamentals and practical aspects. Appl. Phys. Rev. 41, 21305–21302
(2017)
45. Buschow, K.H.J., Sherwood, R.C.: Magnetic properties and hydrogen absorption in rare-earth
intermetallics of the type RMn2 and R6 Mn23 . J. Appl. Phys. 48, 4643–4648 (1977)
46. Holtzberg, F., Gambino, R.J., McGuire, T.R.: New ferromagnetic 5 : 4 compounds in the rare
earth silicon and germanium systems. J. Phys. Chem. Solids. 28, 2283–2289 (1967)
47. Pecharsky, A.O., Gschneidner, K.A., Pecharsky, V.K.: The giant magnetocaloric effect of
optimally prepared Gd5 Si2 Ge2 . J. Appl. Phys. 93, 4722–4728 (2003)
48. Pecharsky, V.K., Gschneidner Jr., K.A.: Giant magnetocaloric effect in Gd5 (Si2 Ge2 ). Phys.
Rev. Lett. 78, 4494–4497 (1997)
49. Pecharsky, V.K., Gschneidner, K.A.: Phase relationships and crystallography in the pseudobi-
nary system Gd5 Si4 -Gd5 Ge4 . J. Alloys Compd. 260, 98–106 (1997)
50. Pecharsky, V.K., Gschneidner, K.A.: Effect of alloying on the giant magnetocaloric effect of
Gd5 (Si2 Ge2 ). J. Magn. Magn. Mater. 167, L179–L184 (1997)
51. Provenzano, V., Shapiro, A.J., Shull, R.D.: Reduction of hysteresis losses in the magnetic
refrigerant Gd5 Ge2 Si2 by the addition of iron. Nature. 429, 853–857 (2004)
52. Han, M., Jiles, D.C., Snyder, J.E., Lograsso, T.A., Schlagel, D.L.: Giant magnetostriction
behavior at the Curie temperature of single crystal Gd5 (Si0.5 Ge0.5 )4 . J. Appl. Phys. 95, 6945–
6947 (2004)
53. Levin, E.M., Pecharsky, V.K., Gschneidner, K.A.: Spontaneous generation of voltage in
Gd5 (Six Ge4-x ) during a first-order phase transition induced by temperature or magnetic field.
Phys. Rev. B. 63, 174110 (2001)
54. Zou, M., Tang, H., Schlagel, D.L., Lograsso, T.A., Gschneidner, K.A., Pecharsky, V.K.:
Spontaneous generation of voltage in single-crystal Gd5 Ge2 Si2 during magnetostructural
phase transformations. J. Appl. Phys. 99, 08B304 (2006)
55. Hadimani, R.L., Silva, J.H.B., Pereira, A.M., Schlagel, D.L., Lograsso, T.A., Ren, Y., Zhang,
X., Jiles, D.C., Araújo, J.P.: Gd5 (Si,Ge)4 thin film displaying large magnetocaloric and strain
effects due to magnetostructural transition. Appl. Phys. Lett. 106, 032402 (2015)
56. Fang, Y.K., Yeh, C.C., Chang, C.W., Chang, W.C., Zhu, M.G., Li, W.: Large low-field
magnetocaloric effect in MnCo0.95 Ge1.14 alloy. Scr. Mater. 57, 453–456 (2007)
57. Trung, N.T., Biharie, V., Zhang, L., Caron, L., Buschow, K.H.J., Brück, E.: From single- to
double-first-order magnetic phase transition in magnetocaloric Mn1−x Crx CoGe compounds.
Appl. Phys. Lett. 96, 162507 (2010)
58. Samanta, T., Dubenko, I., Quetz, A., Stadler, S., Ali, N.: Giant magnetocaloric effects near
room temperature in Mn1 − x Cux CoGe. Appl. Phys. Lett. 101, 242405 (2012)
59. Choudhury, D., Suzuki, T., Tokura, Y., Taguchi, Y.: Tuning structural instability toward
enhanced magnetocaloric effect around room temperature in MnCo1-x Znx Ge. Sci. Rep. 4,
7544 (2014)
60. Yuce, S., Bruno, N.M., Emre, B., Karaman, I.: Accessibility investigation of large magnetic
entropy change in CoMn1-x Fex Ge. J. Appl. Phys. 119, 0–6 (2016)
61. Liu, J., Skokov, K., Gutfleisch, O.: Magnetostructural transition and adiabatic temperature
change in Mn-Co-Ge magnetic refrigerants. Scr. Mater. 66, 642–645 (2012)
62. Tegus, O., Brück, E., Buschow, K.H.J., de Boer, F.R.: Transition-metal-based magnetic
refrigerants for room-temperature applications. Nature. 415, 150–152 (2002)
63. Dung, N.H., Ou, Z.Q., Caron, L., Zhang, L., Thanh, D.T.C., de Wijs, G.A., de Groot, R.A.,
Buschow, K.H.J., Brück, E.: Mixed magnetism for refrigeration and energy conversion. Adv.
Energy Mater. 1, 1215–1219 (2011)
64. Andersson, Y., Rundqvist, S., Beckman, O., Lundgren, L., Nordblad, P.: Properties of
Fe2 P crystals prepared from a liquid copper medium. Phys. Status Solidi. 49, K153–K156
(1978)
30 Magnetocaloric Materials and Applications 1521

65. Wäppling, R., Häggström, L., Ericsson, T., Devanarayanan, S., Karlsson, E., Carlsson, B.,
Rundqvist, S.: First order magnetic transition, magnetic structure, and vacancy distribution in
Fe2 P. J. Solid State Chem. 13, 258–271 (1975)
66. Gercsi, Z., Delczeg-Czirjak, E.K., Vitos, L., Wills, A.S., Daoud-Aladine, A., Sandeman, K.G.:
Magnetoelastic effects in doped Fe2 P. Phys. Rev. B. 88, 024417 (2013)
67. Liu, J., Krautz, M., Skokov, K., Woodcock, T.G., Gutfleisch, O.: Systematic study of the
microstructure, entropy change and adiabatic temperature change in optimized La–Fe–Si
alloys. Acta Mater. 59, 3602–3611 (2011)
68. Hu, F., Shen, B., Sun, J., Cheng, Z., Rao, G., Zhang, X.: Influence of negative lattice expansion
and metamagnetic transition on magnetic entropy change in the compound LaFe11.4 Si1.6 ,
Appl. Phys. Lett. 78, 3675–3677 (2001)
69. Fujita, A., Fujieda, S., Fukamichi, K., Mitamura, H., Goto, T.: Itinerant-electron metamag-
netic transition and large magnetovolume effects in La(Fex Si1-x )13 compounds. Phys. Rev. B.
65 014410 (2001)
70. Fujieda, S., Fujita, A., Fukamichi, K., Yamazaki, Y., Iijima, Y.: Giant isotropic magnetostric-
tion of itinerant-electron metamagnetic La(Fe0.88 Si0.12 )13 Hy compounds. Appl. Phys. Lett.
79, 653 (2001)
71. Fujita, A., Fujieda, S., Hasegawa, Y., Fukamichi, K.: Itinerant-electron metamagnetic transi-
tion and large magnetocaloric effects in La(Fex Si1-x )13 compounds and their hydrides. Phys.
Rev. B. 67 104416 (2003)
72. Fujita, A., Fukamichi, K.: Enhancement of isothermal entropy change due to spin fluctuations
in itinerant-electron metamagnetic La(Fe0.88 Si0.12 )13 compound. J. Alloys Compd. 408–412,
62–65 (2006)
73. Gercsi, Z., Fuller, N., Sandeman, K.G., Fujita, A.: Electronic structure, metamagnetism and
thermopower of LaSiFe12 and interstitially doped LaSiFe12 . J. Phys. D. Appl. Phys. 51,
034003 (2018)
74. Zimm, C.: Proceedings of the 1st International Conference on Magnetic Refrigeration at
Room Temperature Science et Technique Du Froid Comptes Rendus. IIR, Paris (2005)
75. Barcza, A., Katter, M., Zellmann, V., Russek, S., Jacobs, S., Zimm, C.: Stability and
magnetocaloric properties of sintered La(Fe,Mn Si)13 Hz alloys. IEEE Trans. Magn. 47, 3391–
3394 (2011)
76. Baumfeld, O.L., Gercsi, Z., Krautz, M., Gutfleisch, O., Sandeman, K.G.: The dynamics of
spontaneous hydrogen segregation in LaFe13−x Six Hy . J. Appl. Phys. 115, 203905 (2014)
77. Fang, W., Yuan-Fu, C., Guang-Jun, W., Ji-Rong, S., Bao-Gen, S.: Large magnetic entropy
change and magnetic properties in La(Fe1-x Mnx )11.7 Si1.3 Hy compounds. Chinese Phys. 12,
911–914 (2003)
78. Krautz, M., Skokov, K., Gottschall, T., Teixeira, C.S., Waske, A., Liu, J., Schultz, L.,
Gutfleisch, O.: Systematic investigation of Mn substituted La(Fe,Si)13 alloys and their
hydrides for room-temperature magnetocaloric application. J. Alloys Compd. 598, 27–32
(2014)
79. Annaorazov, M.P., Asatryan, K.A., Myalikgulyev, G., Nikitin, S.A., Tishin, A.M., Tyurin,
A.L.: Alloys of the Fe-Rh system as a new class of working material for magnetic
refrigerators. Cryogenics (Guildf). 32, 867–872 (1992)
80. Manekar, M., Roy, S.B.: Reproducible room temperature giant magnetocaloric effect in Fe–
Rh. J. Phys. D. Appl. Phys. 41, 192004 (2008)
81. Stern-Taulats, E., Castán, T., Planes, A., Lewis, L.H., Barua, R., Pramanick, S., Majumdar, S.,
Mañosa, L.: Giant multicaloric response of bulk Fe49 Rh51 . Phys. Rev. B. 95, 104424 (2017)
82. Nikitin, S.A., Myalikgulyev, G., Tishin, A.M., Annaorazov, M.P., Asatryan, K.A., Tyurin,
A.L.: The magnetocaloric effect in Fe49 Rh51 compound. Phys. Lett. A. 148, 363–366 (1990)
83. Annaorazov, M.P., Nikitin, S.A., Tyurin, A.L., Asatryan, K.A., Dovletov, A.K.: Anomalously
high entropy change in FeRh alloy. J. Appl. Phys. 79, 1689–1695 (1996)
84. Stern-Taulats, E., Planes, A., Lloveras, P., Barrio, M., Tamarit, J.L., Pramanick, S., Majumdar,
S., Frontera, C., Mañosa, L.: Barocaloric and magnetocaloric effects in Fe49 Rh51 . Phys. Rev.
B - Condens. Matter Mater. Phys. 89, 1–8 (2014)
1522 K. G. Sandeman and S. Takei

85. Chirkova, A., Skokov, K.P., Schultz, L., Baranov, N.V., Gutfleisch, O., Woodcock, T.G.: Giant
adiabatic temperature change in FeRh alloys evidenced by direct measurements under cyclic
conditions. Acta Mater. 106, 15–21 (2016)
86. Wood, M.E., Potter, W.H.: General analysis of magnetic refrigeration and its optimization
using a new concept: maximization of refrigerant capacity. Cryogenics (Guildf). 25, 667–683
(1985)
87. Zverev, V.I., Tishin, A.M., Kuz’min, M.D.: The maximum possible magnetocaloric T effect.
J. Appl. Phys. 107, 043907 (2010)
88. Sandeman, K.G.: Magnetocaloric materials: the search for new systems. Scr. Mater. 67, 566–
571 (2012)
89. Planes, A., Mañosa, L., Acet, M.: Magnetocaloric effect and its relation to shape-memory
properties in ferromagnetic Heusler alloys. J. Phys. Condens. Matter. 21, 233201 (2009)
90. Pareti, L., Solzi, M., Albertini, F., Paoluzi, A.: Giant entropy change at the co-occurrence
of structural and magnetic transitions in the Ni2.19 Mn0.81 Ga Heusler alloy. Eur. Phys. J. B -
Condens. Matter. 32, 303–307 (2003)
91. Krenke, T., Duman, E., Acet, M., Wassermann, E.F., Moya, X., Mañosa, L., Planes, A.:
Inverse magnetocaloric effect in ferromagnetic Ni–Mn–Sn alloys. Nat. Mater. 4, 450–454
(2005)
92. Chen, X., Srivastava, V., Dabade, V., James, R.D.: Study of the cofactor conditions: conditions
of supercompatibility between phases. J. Mech. Phys. Solids. 61, 2566–2587 (2013)
93. Song, Y., Chen, X., Dabade, V., Shield, T.W., James, R.D.: Enhanced reversibility and unusual
microstructure of a phase-transforming material. Nature. 502, 85–88 (2013)
94. Chluba, C., Ge, W., Lima de Miranda, R., Strobel, J., Kienle, L., Quandt, E., Wuttig, M.:
Shape memory alloys. Ultralow-fatigue shape memory alloy films. Science. 348, 1004–1007
(2015)
95. Dutta, B., Çakır, A., Giacobbe, C., Al-Zubi, A., Hickel, T., Acet, M., Neugebauer, J.: Ab initio
prediction of martensitic and intermartensitic phase boundaries in Ni-Mn-Ga. Phys. Rev. Lett.
116, 025503 (2016)
96. Hamer, J.B.A., Daou, R., Özcan, S., Mathur, N.D., Fray, D.J., Sandeman, K.G.: Phase diagram
and magnetocaloric effect of alloys. J. Magn. Magn. Mater. 321, 3535–3540 (2009)
97. Liu, E., Wang, W., Feng, L., Zhu, W., Li, G., Chen, J., Zhang, H., Wu, G., Jiang, C., Xu,
H., de Boer, F.: Stable magnetostructural coupling with tunable magnetoresponsive effects in
hexagonal ferromagnets. Nat. Commun. 3, 873 (2012)
98. Yu, B., Liu, M., Egolf, P.W., Kitanovski, A.: A review of magnetic refrigerator and heat pump
prototypes built before the year 2010. Int. J. Refrig. 33, 1029–1060 (2010)
99. Guillou, F., Porcari, G., Yibole, H., van Dijk, N., Brück, E.: Taming the first-order transition
in giant magnetocaloric materials. Adv. Mater. 2671–5(2615), 26 (2014)
100. Lyubina, J., Schäfer, R., Martin, N., Schultz, L., Gutfleisch, O.: Novel design of La(Fe,Si)13
alloys towards high magnetic refrigeration performance. Adv. Mater. 22, 3735–3739 (2010)
101. Brück, E., Yibole, H., Zhang, L.: A universal metric for ferroic energy materials. Philos.
Trans. R. Soc. A Math. Phys. Eng. Sci. 374, 20150303 (2016)
102. Morrison, K., Sandeman, K.G., Cohen, L.F., Sasso, C.P., Basso, V., Barcza, A., Katter, M.,
Moore, J.D., Skokov, K.P., Gutfleisch, O.: Evaluation of the reliability of the measurement
of key magnetocaloric properties: a round robin study of La(Fe,Si,Mn)Hδ conducted by the
SSEEC consortium of European laboratories. Int. J. Refrig. 35, 1528–1536 (2012)
103. Guillou, F., Yibole, H., Porcari, G., Zhang, L., van Dijk, N.H., Brück, E.: Magnetocaloric
effect, cyclability and coefficient of refrigerant performance in the MnFe(P,Si,B) system. J.
Appl. Phys. 116, 063903 (2014)
104. Wada, H., Tanabe, Y.: Giant magnetocaloric effect of MnAs1−xSbx. Appl. Phys. Lett. 79,
3302–3304 (2001)
105. Tocado, L., Palacios, E., Burriel, R.: Adiabatic measurement of the giant magnetocaloric
effect in MnAs. J. Therm. Anal. Calorim. 84, 213–217 (2006)
106. Stern-Taulats, E., Planes, A., Lloveras, P., Barrio, M., Tamarit, J.-L., Pramanick, S., Majum-
dar, S., Frontera, C., Mañosa, L.: Barocaloric and magnetocaloric effects in Fe49 Rh51 . Phys.
Rev. B. 89, 214105 (2014)
30 Magnetocaloric Materials and Applications 1523

107. Khovaylo, V.V., Skokov, K.P., Koshkid’Ko, Y.S., Koledov, V.V., Shavrov, V.G., Buchelnikov,
V.D., Taskaev, S.V., Miki, H., Takagi, T., Vasiliev, A.N.: Adiabatic temperature change at first-
order magnetic phase transitions: Ni2.19 Mn0.81 Ga as a case study. Phys. Rev. B - Condens.
Matter Mater. Phys. 78, 1–4 (2008)
108. Madireddi, S.C.: Role of Extrinsic Factors in Utilizing the Giant Magnetocaloric Effect on
Materials: Frequency and Time Dependence. Iowa State University, Digital Repository, Ames
(2010)
109. Gottschall, T., Skokov, K.P., Frincu, B., Gutfleisch, O.: Large reversible magnetocaloric effect
in Ni-Mn-In-Co. Appl. Phys. Lett. 106, 021901 (2015)
110. Tesla, N.: US Patent 396,121 (1889)
111. Edison, T.A.: US Patent 380,100 (1888)
112. Edison, T.A.: US Patent 476,983 (1892)
113. Bremer, H.: US Patent 764,518 (1904)
114. Swiss Blue Energy: Swiss Blue Energy English - YouTube, https://www.youtube.com/watch?
v=e-MwcYYScdY (2016)
115. Hsu, C.-J., Sandoval, S.M., Wetzlar, K.P., Carman, G.P.: Thermomagnetic conversion effi-
ciencies for ferromagnetic materials. J. Appl. Phys. 110, 123923 (2011)
116. Gueltig, M., Wendler, F., Ossmer, H., Ohtsuka, M., Miki, H., Takagi, T., Kohl, M.: High-
performance thermomagnetic generators based on Heusler alloy films. Adv. Energy Mater. 7,
1601879 (2017)
117. Almanza, M., Pasko, A., Mazaleyrat, F., LoBue, M.: Numerical study of thermomagnetic
cycle. J. Magn. Magn. Mater. 426, 64–69 (2017)
118. Bauer, G.E.W., MacDonald, A.H., Maekawa, S.: Spin Caloritronics. Solid State Commun.
150, 459–460 (2010)
119. Bauer, G.E.W., Saitoh, E., van Wees, B.J.: Spin caloritronics. Nat. Mater. 11, 391–399
(2012)
120. Yu, H., Brechet, S.D., Ansermet, J.-P.: Spin caloritronics, origin and outlook. Phys. Lett. A.
381, 825–837 (2017)
121. Maekawa, S., Valenzuela, S.O., Saitoh, E., Kimura, T.: Spin Current. OUP, Oxford (2012)
122. Cahaya, A.B., Tretiakov, O.A., Bauer, G.E.W.: Spin Seebeck power conversion. IEEE Trans.
Magn. 51, 1–14 (2015)
123. Uchida, K., Ota, T., Adachi, H., Xiao, J., Nonaka, T., Kajiwara, Y., Bauer, G.E.W., Maekawa,
S., Saitoh, E.: Thermal spin pumping and magnon-phonon-mediated spin-Seebeck effect. J.
Appl. Phys. 111, 103903 (2012)
124. Kikkawa, T., Uchida, K., Shiomi, Y., Qiu, Z., Hou, D., Tian, D., Nakayama, H., Jin, X.-F.,
Saitoh, E.: Longitudinal spin Seebeck effect free from the proximity Nernst effect. Phys. Rev.
Lett. 110, 67207 (2013)
125. Uchida, K., Ishida, M., Kikkawa, T., Kirihara, A., Murakami, T., Saitoh, E.: Longitudinal
spin Seebeck effect: from fundamentals to applications. J. Phys. Condens. Matter. 26, 343202
(2014)
126. Xiao, J., Bauer, G.E.W., Maekawa, S., Brataas, A.: Charge pumping and the colored thermal
voltage noise in spin valves. Phys. Rev. B. 79, 174415 (2009)
127. Kikkawa, T., Uchida, K., Daimon, S., Shiomi, Y., Adachi, H., Qiu, Z., Hou, D., Jin, X.-
F., Maekawa, S., Saitoh, E.: Separation of longitudinal spin Seebeck effect from anomalous
Nernst effect: determination of origin of transverse thermoelectric voltage in metal/insulator
junctions. Phys. Rev. B. 88, 214403 (2013)
128. Kehlberger, A., Ritzmann, U., Hinzke, D., Guo, E.-J., Cramer, J., Jakob, G., Onbasli, M.C.,
Kim, D.H., Ross, C.A., Jungfleisch, M.B., Hillebrands, B., Nowak, U., Kläui, M.: Length
scale of the spin Seebeck effect. Phys. Rev. Lett. 115, 96602 (2015)
129. Agrawal, M., Serga, A.A., Lauer, V., Papaioannou, E.T., Hillebrands, B.: I. Vasyuchka:
microwave-induced spin currents in ferromagnetic-insulator|normal-metal bilayer system.
Appl. Phys. Lett. 105, 92404 (2014)
130. Uchida, K., Adachi, H., Ota, T., Nakayama, H., Maekawa, S., Saitoh, E.: Observation
of longitudinal spin-Seebeck effect in magnetic insulators. Appl. Phys. Lett. 97, 172505
(2010)
1524 K. G. Sandeman and S. Takei

131. Ritzmann, U., Hinzke, D., Kehlberger, A., Guo, E.-J., Kläui, M., Nowak, U.: Magnetic field
control of the spin Seebeck effect. Phys. Rev. B. 92, 174411 (2015)
132. Jin, H., Boona, S.R., Yang, Z., Myers, R.C., Heremans, J.P.: Effect of the magnon dispersion
on the longitudinal spin Seebeck effect in yttrium iron garnets. Phys. Rev. B. 92, 54436 (2015)
133. Schmid, M., Srichandan, S., Meier, D., Kuschel, T., Schmalhorst, J.-M., Vogel, M., Reiss, G.,
Strunk, C., H, C.: Back: transverse spin Seebeck effect versus anomalous and planar Nernst
effects in permalloy thin films. Phys. Rev. Lett. 111, 187201 (2013)
134. Tikhonov, K.S., Sinova, J., Finkel’stein, A.M.: Spectral non-uniform temperature and non-
local heat transfer in the spin Seebeck effect. Nat. Commun. 4, 1945 (2013)
135. Ganzhorn, K., Wimmer, T., Cramer, J., Schlitz, R., Geprägs, S., Jakob, G., Gross, R., Huebl,
H., Kläui, M., Goennenwein, S.T.B.: Temperature dependence of the non-local spin Seebeck
effect in YIG/Pt nanostructures. AIP Adv. 7, 085102 (2017)
136. Weiler, M., Althammer, M., Czeschka, F.D., Huebl, H., Wagner, M.S., Opel, M., Imort, I.-M.,
Reiss, G., Thomas, A., Gross, R., Goennenwein, S.T.B.: Local charge and spin currents in
magnetothermal landscapes. Phys. Rev. Lett. 108, 106602 (2012)
137. Ramos, R., Kikkawa, T., Uchida, K., Adachi, H., Lucas, I., Aguirre, M.H., Algarabel, P.,
Morellón, L., Maekawa, S., Saitoh, E., Ibarra, M.R.: Observation of the spin Seebeck effect
in epitaxial Fe3 O4 thin films. Appl. Phys. Lett. 102, 72413 (2013)
138. Xiao, J., Bauer, G.E.W., Uchida, K., Saitoh, E., Maekawa, S.: Theory of magnon-driven spin
Seebeck effect. Phys. Rev. B. 81, 214418 (2010)
139. Adachi, H., Ohe, J., Takahashi, S., Maekawa, S.: Linear-response theory of spin Seebeck
effect in ferromagnetic insulators. Phys. Rev. B. 83, 94410 (2011)
140. Hoffman, S., Sato, K., Tserkovnyak, Y.: Landau-Lifshitz theory of the longitudinal spin
Seebeck effect. Phys. Rev. B. 88, 64408 (2013)
141. Rezende, S.M., Rodríguez-Suárez, R.L., Cunha, R.O., Rodrigues, A.R., Machado, F.L.A.,
Fonseca Guerra, G.A., Lopez Ortiz, J.C., Azevedo, A.: Magnon spin-current theory for the
longitudinal spin-Seebeck effect. Phys. Rev. B. 89, 14416 (2014)
142. Schreier, M., Kamra, A., Weiler, M., Xiao, J., Bauer, G.E.W., Gross, R., Goennenwein, S.T.B.:
Magnon, phonon, and electron temperature profiles and the spin Seebeck effect in magnetic
insulator/normal metal hybrid structures. Phys. Rev. B. 88, 94410 (2013)
143. Tserkovnyak, Y., Brataas, A., Bauer, G.E.W., Halperin, B.I.: Nonlocal magnetization dynam-
ics in ferromagnetic heterostructures. Rev. Mod. Phys. 77, 1375 (2005)
144. Brataas, A., Bauer, G.E.W., Kelly, P.J.: Non-collinear magnetoelectronics. Phys. Rep. 427,
157–255 (2006)
145. Agrawal, M., Vasyuchka, V.I., Serga, A.A., Kirihara, A., Pirro, P., Langner, T., Jungfleisch,
M.B., Chumak, A.V., Papaioannou, E.T., Hillebrands, B.: Role of bulk-magnon transport in
the temporal evolution of the longitudinal spin-Seebeck effect. Phys. Rev. B. 89, 224414
(2014)
146. Guo, E.-J., Cramer, J., Kehlberger, A., Ferguson, C.A., MacLaren, D.A., Jakob, G., Kläui, M.:
Influence of thickness and interface on the low-temperature enhancement of the spin Seebeck
effect in YIG films. Phys. Rev. X. 6, 31012 (2016)
147. Ritzmann, U., Hinzke, D., Nowak, U.: Propagation of thermally induced magnonic spin
currents. Phys. Rev. B. 89, 24409 (2014)
148. Rezende, S.M., Rodríguez-Suárez, R.L., Azevedo, A.: Theory of the spin Seebeck effect in
antiferromagnets. Phys. Rev. B. 93, 14425 (2016)
149. Seki, S., Ideue, T., Kubota, M., Kozuka, Y., Takagi, R., Nakamura, M., Kaneko, Y., Kawasaki,
M., Tokura, Y.: Thermal generation of spin current in an antiferromagnet. Phys. Rev. Lett. 115,
266601 (2015)
150. Wu, S.M., Zhang, W., KC, A., Borisov, P., Pearson, J.E., Jiang, J.S., Lederman, D., Hoffmann,
A., Bhattacharya, A.: Antiferromagnetic spin Seebeck effect. Phys. Rev. Lett. 116, 97204
(2016)
151. Holanda, J., Maior, D.S., Alves Santos, O., Vilela-Leão, L.H., Mendes, J.B.S., Azevedo, A.,
Rodríguez-Suárez, R.L., Rezende, S.M.: Spin Seebeck effect in the antiferromagnet nickel
oxide at room temperature. Appl. Phys. Lett. 111, 172405 (2017)
30 Magnetocaloric Materials and Applications 1525

152. Ribeiro, P.R.T., Machado, F.L.A., Gamino, M., Azevedo, A., Rezende, S.M.: Spin Seebeck
effect in antiferromagnet nickel oxide in wide ranges of temperature and magnetic field. Phys.
Rev. B. 99, 094432 (2019)
153. Uchida, K., Takahashi, S., Harii, K., Ieda, J., Koshibae, W., Ando, K., Maekawa, S., Saitoh,
E.: Observation of the spin Seebeck effect. Nature. 455, 778–781 (2008)
154. Jaworski, C.M., Yang, J., Mack, S., Awschalom, D.D., Heremans, J.P., Myers, R.C.: Obser-
vation of the spin-Seebeck effect in a ferromagnetic semiconductor. Nat. Mater. 9, 898–903
(2010)
155. Adachi, H., Uchida, K., Saitoh, E., Ohe, J., Takahashi, S., Maekawa, S.: Gigantic enhance-
ment of spin Seebeck effect by phonon drag. Appl. Phys. Lett. 97, 252506 (2010)
156. Jaworski, C.M., Yang, J., Mack, S., Awschalom, D.D., Myers, R.C., Heremans, J.P.: Spin-
Seebeck effect: a phonon driven spin distribution. Phys. Rev. Lett. 106, 186601 (2011)
157. Shan, J., Cornelissen, L.J., Vlietstra, N., Ben Youssef, J., Kuschel, T., Duine, R.A., van Wees,
B.J.: Influence of yttrium iron garnet thickness and heater opacity on the nonlocal transport
of electrically and thermally excited magnons. Phys. Rev. B. 94, 174437 (2016)
158. Sinova, J., Valenzuela, S.O., Wunderlich, J., Back, C.H., Jungwirth, T.: Spin Hall effects. Rev.
Mod. Phys. 87, 1213–1260 (2015)
159. Cheng, S., Xing, Y., Sun, Q., Xie, X.C.: Spin Nernst effect and Nernst effect in two-
dimensional electron systems. Phys. Rev. B. 78, 45302 (2008)
160. Tauber, K., Gradhand, M., Fedorov, D.V., Mertig, I.: Extrinsic spin Nernst effect from first
principles. Phys. Rev. Lett. 109, 26601 (2012)
161. Wimmer, S., Ködderitzsch, D., Chadova, K., Ebert, H.: First-principles linear response
description of the spin Nernst effect. Phys. Rev. B. 88, 201108 (2013)
162. Meyer, S., Chen, Y.-T., Wimmer, S., Althammer, M., Geprägs, S., Huebl, H., Ködderitzsch,
D., Ebert, H., Bauer, G.E.W., Gross, R., Goennenwein, S.T.B.: Observation of the spin Nernst
effect. Nat. Mater. 16, 977 (2017)
163. Sheng, P., Sakuraba, Y., Takahashi, S., Mitani, S., Hayashi, M.: Signatures of the spin Nernst
effect in tungsten. Sci. Adv. 3, e1701503 (2017)
164. Murakami, S.: Dissipationless quantum spin current at room temperature. Science (80-. ).
301, 1348–1351 (2003)
165. Sinova, J., Culcer, D., Niu, Q., Sinitsyn, N.A., Jungwirth, T., MacDonald, A.H.: Universal
intrinsic spin Hall effect. Phys. Rev. Lett. 92, 126603 (2004)
166. Kovalev, A.A., Zyuzin, V.: Spin torque and Nernst effects in Dzyaloshinskii-Moriya ferro-
magnets. Phys. Rev. B. 93, 161106 (2016)
167. Kim, S.K., Ochoa, H., Zarzuela, R., Tserkovnyak, Y.: Realization of the Haldane-Kane-Mele
model in a system of localized spins. Phys. Rev. Lett. 117, 227201 (2016)
168. Zyuzin, V.A., Kovalev, A.A.: Magnon spin Nernst effect in antiferromagnets. Phys. Rev. Lett.
117, 217203 (2016)
169. Cheng, R., Okamoto, S., Xiao, D.: Spin Nernst effect of magnons in collinear antiferromag-
nets. Phys. Rev. Lett. 117, 217202 (2016)
170. Kato, Y.K., Myers, R.C., Gossard, A.C., Awschalom, D.D.: Observation of the spin Hall effect
in semiconductors. Science (80-. ). 306, 1910–1913 (2004)
171. Dyakonov, M.I., Perel, V.I.: Current-induced spin orientation of electrons in semiconductors.
Phys. Lett. A. 35, 459–460 (1971)
172. Hirsch, J.E.: Spin Hall effect. Phys. Rev. Lett. 83, 1834–1837 (1999)
173. Wunderlich, J., Kaestner, B., Sinova, J., Jungwirth, T.: Experimental observation of the spin-
Hall effect in a two-dimensional spin-orbit coupled semiconductor system. Phys. Rev. Lett.
94, 47204 (2005)
174. Saitoh, E., Ueda, M., Miyajima, H., Tatara, G.: Conversion of spin current into charge current
at room temperature: inverse spin-Hall effect. Appl. Phys. Lett. 88, 182509 (2006)
175. Valenzuela, S.O., Tinkham, M.: Direct electronic measurement of the spin Hall effect. Nature.
442, 176 (2006)
176. Kitanovski, A.: Energy applications of magnetocaloric materials. Adv. Energy Mater. 10,
1903741 (2020)
1526 K. G. Sandeman and S. Takei

177. Barclay, J., Brooks, K., Cui J., Holladay J., Meinhardt K„ Polikarpov E., Thomsen E.: Propane
liquefaction with an active magnetic regenerative liquefier. Cryogenics. 100, 69–76 (2019)
178. Kirol, L.D., Dacus, M.W.: Rotary recuperative magnetic heat pump. In: Cryogenic Engineer-
ing Conference, St. Charles, pp. 757–765 (1987)
179. Kirol, L.D.: Rotary magnetic heat pump. US patent 4727722 (1988)
180. Tan, X., Chai, P., Thompson, C.M., Shatruk, M.: Magnetocaloric effect in AlFe2B2: toward
magnetic refrigerants from earth-abundant elements. J. Am. Chem. Soc. 135, 9553–9557
(2013)
181. Lejeune, B., Schlagel, D.L., Jensen, B., Lograsso, T.A., Kramer, M.J., Lewis, L.: Effects of
Al and Fe solubility on the magnetofunctional properties of AlFe2B2. Phys. Rev. Mater. 3,
094411 (2019)
182. Oey, Y.M., Bocarsly, J.D., Mann, D., Levin, E.E., Shatruk, M., Seshadri, R.: Structural
changes upon magnetic ordering in magnetocaloric AlFe2B2. Appl. Phys. Lett. 116, 212403
(2020)
183. Lewis, L.H., Marrows, C.H., Langridge, S.: Coupled magnetic, structural, and electronic
phase transitions in FeRh. J. Phys. D: Appl. Phys. 49, 323002 (2016)
184. Dzekan, D., Waske, A., Nielsch, K., Fähler, S.: Efficient and affordable thermomagnetic
materials for harvesting low grade waste heat. APL Mater. 9, 011105 (2021)

Karl G. Sandeman received his PhD in Condensed Matter


Theory from the University of Cambridge in 2003. He was subse-
quently awarded a Royal Society University Research Fellowship
to work at Cambridge then Imperial College London, where he
became a Senior Lecturer in Physics. His research at CUNY
uses experimental and computational tools to examine solid-state
phase transitions, often around room temperature.

So Takei is an assistant professor in the Physics Department at


CUNY Queens College. He earned his Ph.D. in 2008 from the
University of Toronto and worked at the Max Planck Institute
for Solid State Research, University of Maryland and UC Los
Angeles. His research area is in theoretical condensed matter
physics with expertise in spintronics and quantum magnetism.
Magnetic Sensors
31
Myriam Pannetier-Lecoeur and Claude Fermon

Contents
Magnetic Sensing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1528
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1528
Relevant Parameters for Choice of a Magnetic Sensor Source Signal Parameters . . . . . . 1528
Field or Flux Sensing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1529
Sensor Types . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1530
Flux Sensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1530
Field Sensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1533
Noise . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1541
Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1541
Noise Sources [31] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1542
Detection Schemes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1542
Magnetometer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1542
Gradiometer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1543
Closed Loop . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1544
Bridge . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1544
Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1544
Field/Flux Sensing and Flux Transformers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1544
Current Sensing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1545
Position Sensing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1545
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1548

Abstract

Here the fundamentals of magnetic sensing are introduced and several types
of magnetic flux and field sensors are presented. Flux sensors are inductive

M. Pannetier-Lecoeur () · C. Fermon


Service de Physique de l’Etat Condensé, DRF/IRAMIS/SPEC CNRS UMR 3680 CEA Saclay,
Gif sur Yvette, France
e-mail: myriam.lecoeur@cea.fr; claude.fermon@cea.fr

© Springer Nature Switzerland AG 2021 1527


J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_30
1528 M. Pannetier-Lecoeur and C. Fermon

coils, fluxgates, and SQUIDs. Field sensors depend on Hall effect, anisotropic
magnetoresistance, spin electronics (giant magnetoresistance and tunnel mag-
netoresistance), optical pumping, giant magnetoimpedance, and magnetoelectric
effects. Some specific readout schemes are discussed. The noise aspect is dis-
cussed and comparative tables are provided of the various sensors including their
sensitivity, field range, minimal detectable noise, advantages, and drawbacks.

Magnetic Sensing

Introduction

A sensor is an entity which converts a physical quantity into another. Like our
sensory system, which changes the physical inputs (light and color, sound . . . ) into
electrical signals which can be conveyed through the nervous system towards var-
ious brain areas, a manufactured sensor will change a physical quantity containing
relevant information to another which can be easily interfaced and treated, most
often an electrical signal. Sensing the magnetic field H is a powerful approach as it
does not require physical contact with the source of the signal. Magnetic sensors
give an output signal containing information on B, which can be related to its
magnitude (scalar magnetometer) or to its components (vector magnetometer).
The magnetic field H is given in A/m or Oe. The magnetic induction B is given
in tesla T (or gauss G). As for magnetic field detection (in a free space), they are
proportional, magnetic sensor performance like sensitivity, field equivalent noise,
and response are given in tesla which is the most convenient unit.
Magnetic fields can be produced by many sources, either from a magnetic object
(from nanoscale magnetic particles to celestial objects such as planets or stars) or
induced by an electrical current. As the variety of magnetic signals which can be
explored is broad, many parameters influence the choice of a sensor for a particular
application (Fig. 1).

Relevant Parameters for Choice of a Magnetic Sensor Source Signal


Parameters

The nature of the source defines the characteristics of the magnetic signal emitted
and hence drives the choice of sensor. For instance, the source size and shape (point-
like source, dipolar or distributed and homogeneous sources) will define the field
shape and extension. The possible distance between the sources and the measure-
ment site will lead to the field strength. From these pieces of information, the sensor
type, size, and configuration will be chosen. Additional important parameters are the
frequency of the emitted signal (DC up to GHz or THz frequencies), the temporal
31 Magnetic Sensors 1529

Fig. 1 Schematic of a magnetic measurement: the source characteristics, the distance from the
sensor to the source, the allowable size, and the environmental conditions are among the parameters
that drive the choice of sensor

response (continuous or transient). These parameters will fix the sensor dynamic
range and required bandwidth.

Sensor Characteristics
Magnetic sensor characteristics are complex and strongly dependent on the physical
principle underlying the sensor operation [1]. Sensitivity and field equivalent noise
are key parameters, but dynamics, frequency, and temperature range are also to be
considered. Sensors can also be operated as vector magnetometers, measuring one,
two, or three components, or in scalar mode, where only field strength is sensed.
Depending on their principle, sensors can be absolute or relative magnetometers.
Sensors can also be defined in terms of energy resolution versus volume [2].

External Parameters
The choice for a particular sensor will be based upon the conditions under which
the sensor has to be operated (temperature range, electromagnetic perturbations,
allowable size and weight, price . . . ).

Field or Flux Sensing

Some sensors are designed for flux sensing where H is integrated over a finite
surface area defined by the sensor shape. If size is not a limitation and if the source
1530 M. Pannetier-Lecoeur and C. Fermon

is distant, flux sensors can be appropriate, whereas for local measurements or when
miniaturization is required, field sensors will be a better option.
Flux sensors can be either inductive, such as coils, and dedicated to high
frequency (≥10 kHz) measurements, or sensitive to a static or low frequency flux
such as SQUIDS and fluxgates. Main types of flux and field sensors are detailed
below.

Sensor Types

Flux Sensors

In this section, sensors which measure the magnetic flux Φ are addressed. Such
sensors can be extremely field sensitive when their size is large, hence Φ is large
as well. Coils, which are the most common flux sensors, are very sensitive for
high frequency signals, whereas fluxgates and SQUIDs address low frequency or
dc signals.

Coils and Inductive Sensors

Principle
When an open metallic coil is submitted to an external varying field, a voltage
proportional to the time derivative of the magnetic flux appears at the ends of the
coil. This simple sensor is still used extensively as a high frequency receiver. The
intrinsic level of noise of a coil, only given by the thermal noise of the residual
resistance, is very low and the sensitivity is directly proportional to the frequency.
Hence at high frequencies, coils can easily reach sub-femtotesla detection [1, 3, 4].
From the simple coil, two main alternative sensors have been developed. The first
consists of placing the coil in a resonant circuit by adding a capacitance (Fig. 2a).
This is very useful as a radio receiver or as an NMR spectrometer. The output
voltage is:

d
V = −Q (1)
dt

for a resonant coil where Q is the quality factor of the coil and it depends on the
resonant circuit characteristics and on the frequency (Fig. 3). The second consists
of a ferrite core with a high permeability for detecting low frequencies (Fig. 2b).
Search coils used for petroleum research or long wave radio receivers use this latter
method.

Main Applications
Receivers, NMR, and MRI spectrometers are using tuned coils for frequency-
specific detection. Low-frequency detection is performed with search coils compris-
ing a large ferrite core. Proximity inductive sensors and nondestructive evaluation
31 Magnetic Sensors 1531

Fig. 2 Coil sensors: (a) Resonant coil circuit; (b) Search coil, where a ferrite is placed within the
winding. The increased μr leads to an effective increase of sensitivity, equivalent to an enlarged
pickup surface

Fig. 3 Output voltage amplitude as a function of frequency for a simple coil (grey line) and for a
resonant coil (dashed line) tuned at 5 kHz

use the same approach: a small coil creates a RF field which induces an eddy current
in the metallic target and the same coil (or another one) is used to detect the field
created by these currents. This can also be seen as a change of inductance of the
emitting coil.

Fluxgates

Principle
Fluxgates are also designed with a soft magnetic material core surrounded by a
coil. However, the working principle is quite different. The coil is used to apply
an alternating magnetic field in the kHz range (at a frequency f ) able to saturate
the core. A second sensing coil is placed around the system (Fig. 4). In absence of
1532 M. Pannetier-Lecoeur and C. Fermon

Fig. 4 Fluxgate principle: A soft magnetic core, often in a ring shape is symmetrically excited
by an ac coil creating field large enough to saturate the core. The material hence experiences a
complete hysteresis loop at each cycle. Due to the symmetry, in absence of external field, there is
no magnetization created by the ring. When an external magnetic field is applied, the left and right
hysteresis loops are shifted in opposite directions and a net magnetization appears, detected by the
reception coil. The signal at 2f is proportional to the magnitude of the external field

external field, the response of the core is symmetrical and there is no signal in the
sensing coil. When an external field is applied, the response of the core is no longer
symmetrical and a response at 2f, the double of the applied frequency, is detected
[1, 5].
Fluxgates have been extensively developed with small, μm (micro-fluxgates) or
cm sizes and are able to detect fields in the range of picoteslas to nanoteslas. Unlike
classical coils, they are sensitive to dc fields and have a reasonably flat response up
to several kHz [6].

Main Applications
The main application of fluxgates is to detect fields in the range of 100 pT to
1 μT. This applies mainly for earth and planetary magnetic field measurements,
compasses or for laboratory instrumentation. Due to their low 1/f noise, fluxgates are
often used as references for active magnetic shielding of rooms used for biomagnetic
measurements.

SQUIDs

Principle
Superconducting quantum interferences devices (SQUIDs) [7] are based on two
main properties of superconductivity; the first is the quantization of the flux and
the second is the tunneling transport of Cooper pairs through a weak link in a
superconducting film, also called a Josephson Junction (JJ). SQUIDs comprise a
31 Magnetic Sensors 1533

Fig. 5 Principle of SQUID measurement. The field of interest is captured in a pickup coil which is
coupled to a second loop (primary coil) directly coupled to the SQUID. Such flux transformers are
made of superconducting material, like the SQUID, which is a superconducting loop interrupted
by Josephson Junctions, which are tunnel junctions or weakly superconducting links. Both flux
transformer and SQUIDs require a cryogenic environment to operate. Usually, SQUIDs are used
with a flux-locked loop scheme (see Sect. “Closed Loop”), which allows operating at the most
sensitive part of the output voltage

superconducting loop containing one (dc-SQUIDS) or two (dc-SQUIDs) junctions,


the junction being either an insulator or a normal metal.
The output voltage of a SQUID exhibits a modulation with a period correspond-
ing to one flux quantum Φ 0 (= 2.07 10−15 Wb) on the loop. In dc-SQUIDS, the
two JJs interference gives also rise to a slower oscillation, the overall response
being analogous to Fraunhofer patterns in optical diffraction through two slits.
Dc-SQUIDS are biased with constant current whereas rf-SQUIDS are biased with
oscillating rf current. SQUIDs are usually coupled to a superconducting flux
transformer comprising a larger pickup coil and a second coil coupled to the SQUID
loop (Fig. 5). To operate, SQUIDs need to be cooled below the critical temperature
of the superconducting loop, with liquid nitrogen for high-Tc SQUIDs and with
liquid helium for low-Tc SQUIDs (usually made of Niobium). Frequency range of
operation is very large (dc to GHZ) and sensitivity is among the best for static or
low frequency signals, with extremely low noise (especially for low-Tc SQUIDs).

Main Applications
SQUIDs are used in metrology, laboratory magnetometry, explosives detection, or
geophysical measurements and mapping [8] with the main commercial application
being magnetoencephalography [9]. Nevertheless, the need of cryogenic cooling
limits their practical applicability to detect extremely weak low frequency signals,
such as biomagnetic ones (heart or brain magnetic activity for instance) [8]
(Table 1).

Field Sensors

Field sensors directly measure vector components or the absolute value of H. Here
we present the main types of magnetic field sensors.
1534 M. Pannetier-Lecoeur and C. Fermon

Table 1 Flux sensors main characteristics


Coil Search coil Fluxgate SQUID
Detection axis Perpendicular Axial Axial Perpendicular
Temperature a −273/+600 ◦ C −50/+200 ◦ Cb −50/+200 ◦ C b −273/−196 ◦ Cc

range
Typical field 1fT–10 T 1fT–10 mT 5 pT–100 μT
range 1 fT–100 μT
Frequency range ac >50 Hz ac >5 mHz [10] dc-5 kHz dc-100 kHz
Linearity Excellent Excellent Good Requires feedback
Size 0.1 mm–1 m 0.1 mm–1 m 0.1 mm–5 cm 0.1 mm–1 cm
Material Metal Ferrite core Ferromagnet Metal/oxide
Technology Bulk Bulk Bulk Thin film
a Sensor operating temperature range
b Limited by the material permeability range
c Given by the superconducting transition temperature

Hall Effect Sensors

Principle
Discovered by E.H. Hall in 1879 [11], the Hall effect describes charge carrier
deflection in a metal or a semiconductor under the Lorentz force towards the edge
of a strip, leading to a transverse voltage, directly proportional to H (Fig. 6).
In first approximation, the Hall voltage is given by:

Rh
VH = BI. (2)
t

with t the thickness of the slab and where the Hall resistance is defined as

1
Rh = . (3)
q.nc

q and nc being respectively the charge and the carrier density.

Fig. 6 Hall effect principle:


charge carriers (here
electrons) are deflected by a
field applied out of the plane
of a semiconducting material
towards one edge of the slab,
generating a voltage
proportional to H
31 Magnetic Sensors 1535

Main Applications
A Hall sensor can be a simple strip connected with two contacts to inject the
current and two transverse electrodes reading the output voltage. Best materials for
Hall sensors are semiconductors due to their large Rh . Because of its simplicity
and compatibility with microelectronics, Hall effect is the main physical principle
widely exploited in consumer and automotive sensor applications.

AMR Sensors

Principle
Anisotropic magnetoresistance (AMR) appears in a magnetic material as a resistiv-
ity change when the relative orientation of the current is altered in respect to the
magnetization direction.
Discovered by W. Thomson in 1857 [12], it can be expressed by

ρ = ρperp + ρ cos2 θ (4)

with

ρ = ρpar − ρperp (5)

where ρ par is the resistivity when the current flows parallel to M, i.e., for θ = 0,
and ρ perp the resistivity when the current flows perpendicular to M, i.e., for θ = 90◦
(Fig. 7).

Fig. 7 AMR principle: The resistance of a ferromagnetic slab in which a current flow depends
on the field H applied in the plane of the slab, and varies as cos (θ)2 ; θ being the angle between
the current and the magnetization (see in Fig. 8a). To linearize the sensor response, the best way
is to set this angle to 45◦ , for instance, by adding some conducting lines (shown here in gray) to
force the current flow at 45◦ to the magnetization. This configuration is called “barber pole.” The
resistance variation as a function of field is given in Fig. 8b
1536 M. Pannetier-Lecoeur and C. Fermon

Fig. 8 (a) Resistance variation of the AMR as a function of the angle between the applied field
H and the current. (b) Resistance as a function of a field aligned at 45◦ of the current flow. The
response is linear at low field

Main Applications
AMR thin films were once used to replace the inductive coils in computers read
heads, thanks to their good sensitivity and to their miniaturization. They are used
for linear sensors and current sensors.

Spin Electronic Sensors (GMR, TMR)


Spin electronic sensors rely on the magnetoresistance properties arising in stacks of
magnetic materials with thicknesses that allow the spin coherence to be conserved.
Their development has been strongly linked to that of thin film technology over the
last 30 years.

Principle
Giant magnetoresistance appears in ferromagnetic materials which are magnetically
coupled through a very thin metallic layer [13]. First observed in Fe/Cr multilayers
[14], it has been extended to more elaborate stacks, such as the spin valve proposed
by Dieny in 1991 [15] which includes a reference layer by coupling a ferromagnet
to an antiferromagnet and a free layer comprising a ferromagnet; both pinned
and free layers are separated by a nonmagnetic metal, such as copper. In tunnel
magnetoresistance (TMR), the metallic spacer is replaced by an insulating layer
through which the conduction electrons from the pinned (or the free) layer tunnel
to the free (or pinned layer). In GMR sensors, magnetoresistance ratios as high as
20% can be obtained, whereas in TMR sensor with an MgO spacer, it can exceed
200% at room temperature [16, 17]. The physical origin of these effects is the spin
polarization of conduction electrons in metallic magnetic layers, hence in a spin
valve structure, for example, a first magnetic layer acts as a spin polarizer whereas
the second magnetic layer (usually made of soft ferromagnet) operates as a spin
analyzer. The whole structure filters electrons according to the relative orientation
of the two layer’s magnetization.
31 Magnetic Sensors 1537

Fig. 9 GMR in angle sensing mode. (a) The structure comprises a pinned layer (PL) – here shown
pointing perpendicularly towards the bottom – whose magnetization is set by the choice of material
and deposition process conditions, and a free layer (FL), made of layers of ferromagnetic material
with a very low Hc , such as a NiFe/CoFe bilayer. The two sets of layers, PL and FL are separated
by a spacer of nonmagnetic material, conducting for GMR (copper being the most commonly used)
and insulating (such as MgO or Al2 O3 ) for TMR. When a field H > Hsat is applied in the plane of
the sensor, it aligns the FL layer and the resistance of the sensor is proportional to the cosine of θ,
the angle between the PL and the FL. (b) Output voltage of the GMR in the angle sensing mode

In a spin valve, which is the most commonly used stack for magnetic sensing,
the output voltage is given by:

 
Rap − Rp
V = R0 + cos θ I, (6)
2

Rap (resp. Rp ) is the resistance when the pinned and free layer are antiparallel
(resp. parallel), R0 the mean value, and θ the angle between the pinned and the free
layer (see also Fig. 9).
The sensor response can be linearized around H = 0 by means of an additional
in-plane magnetization (Fig. 10) to obtain an output voltage proportional to H .

Main Applications
Devices based on GMR or TMR have replaced AMR technology in hard disk read
heads, which has been the largest application of spin electronics. Now magnetic
random access memories (MRAMs) based on TMR are marketed, and sensing
applications in the area of automotive or other large-scale application (mobile
phones . . . ) are developing. GMR and TMR can be integrated to CMOS in a back-
end process, allowing small size integrated devices.
1538 M. Pannetier-Lecoeur and C. Fermon

Fig. 10 GMR in field sensing mode. (a) On the same type of GMR stack as shown in Fig. 9a,
the addition of a magnetization M0 in plane and perpendicular to the FL leads to a voltage change
proportional to the angle between the applied field and M0 , therefore proportional to the magnitude
of H . (b) Corresponding output voltage of the GMR sensor

Optically Pumped Magnetometers

Principle
First demonstrated in the 1960s [18, 19] on a principle shown in 1957 [20], optical
magnetometers use alkali atoms which are spin polarized by resonant light. Under
magnetic field, electron spins precess and the light transmitted through the atomic
vapor is reduced. The resolution of these magnetometers is limited by the spin
exchange relaxation which brings decoherence, but by increasing the density of
the atoms in the vapor in the cell (by heating), field detection levels down to

1fT/ Hz have been demonstrated in so-called SERF (for spin exchange relaxation
free) optical magnetometers [21] (Fig. 11).
Sensitivity to magnetic field for a time of measurement tM is given by [22]:

1 
δB = √ (7)
gμ0 Nτ tM

g being the Landé factor, μ0 the Bohr magneton,  the Planck constant, τ the
coherence time, and N the number of atoms.

Main Applications
Thanks to their high sensitivity and low noise, optical magnetometers are used
for geomagnetic field mapping, biomagnetic signal detection, material science, or
NMR and MRI. They do not require cryogenics and can be made in relatively small
size/chip-scale devices [23]. The main limitation is the dynamics, which is limited
in field range for the most sensitive magnetometer, and requires a heavy magnetic
shielding to ensure sufficient screening of electromagnetic perturbations.
31 Magnetic Sensors 1539

Fig. 11 SERF optical magnetometer principle: The alkali vapor contained in the gas cell is
polarized by circularly polarized light (pump laser). The probe laser light, partially absorbed when
a field is applied to the alkali atoms, is detected by an array of photodiodes

Nitrogen Vacancies Centers

Principle
Nitrogen vacancies (NV) centers are one of the point defects in diamond. They are
particularly useful as individual center can be detected by photoluminescence. NV
fluorescence is spin-state dependent. Hence a microwave sent at 2.88 GHz induces
a change in the fluorescence intensity. In presence of an external field, the frequency
of absorption is split in two lines with a separation given by the g factor [53, 52]. By
measuring the splitting, the field can be deduced. The present detectivity of a single
√ √
center is in the range of 1 μT/ Hz and for an assembly of centers, sub-nT/ Hz
detectivity are achieved.

Main Applications
NV centers are very attractive for two applications: a very local detection of
magnetic fields with resolution achieving today 50 nm [51, 50] and magnetic
imaging at larger scale. More recently NV centers have been also used for nanometer
scale NMR spectroscopy [49]. NV centers have also a long coherence time which
make them very attractive also or quantum storage and quantum computing.

Giant Magnetoimpedance

Principle
Discovered in 1994 [24, 25], giant magnetoimpedance (GMI) is based on a physical
phenomenon totally different from the spin electronic-based sensors. When a
1540 M. Pannetier-Lecoeur and C. Fermon

metallic wire is subject to an ac voltage, due to the skin effect, the current flows
mainly at the surface of the wire. If this wire is made of magnetic material, its
permeability has an impact on the conductivity and hence the impedance of the wire
will be dependent on its static magnetic configuration. An external dc magnetic field,
by changing this configuration, induces a change in the measured impedance. This
effect is highly nonlinear but can be rather large and in a closed loop design, highly
sensitive devices have been proposed [26, 27, 28, 29].

Main Applications
Presently, GMI is investigated for applications such as current sensors, position
sensors, or for biochip applications but its nonintegrability and the necessity of
having high ac frequency and data treatment limit its use to specific high-sensitivity
applications.

Magnetoelectric Sensors

Principle
The magnetoelectric response is the appearance of an electric polarization P
upon applying a magnetic field H and/or the appearance of a magnetization M
upon applying an electric field E (inverse magnetoelectric effect). This effect is
extensively studied in multiferroic components for applications such as magnetic
sensors. However, based on strain-induced by magnetostriction, magnetoelectric
sensors can also be built. If a piezoelectric material such as PZT (Lead Zirconate
Titanate) is strongly coupled to a soft magnetic material, a voltage applied to the
PZT creates a strain in the material which can modify the direction or the amplitude
of the magnetization. Inversely, if a magnetic field is applied on a magnetic material,
due to magnetostriction, a strain is created in the piezoelectric material inducing a
voltage (Fig. 12). The interest of this kind of sensor is that it is passive, as a voltage
is directly created by a field change with no external power supply (Tables 2 and 3).

Fig. 12 Example of a
magnetoelectric device. A
soft magnetic material like Ni
with a relatively large
magnetostriction is coupled
to a piezoelectric material.
When a magnetic field is
applied, the magnetostrictive
material is creating a strain on
the piezoelectric material
leading to the appearance of a
voltage
31 Magnetic Sensors 1541

Table 2 Field sensors main characteristics – 1


Hall AMR GMR/TMR
Detection axis Perp. In plane In plane
Temperature range −200/+150 ◦ C −273/+200 ◦ C −273/+200 ◦ C
Typical field range 1 μT–10 T 1 nT–1 mT 100 pT–10 mT
Frequency range dc-1 MHz dc-10 MHz dc-GHz
Linearity Good Limited range of H Limited range of H
Size μm-mm μm-mm μm-mm
Material Semiconductor Ferromagnet Multilayers
Technology Thin film Thin film Thin film

Table 3 Field sensors main characteristics – 2


GMI Magnetoelectric Optical pumpinga
Detection axis Axial In plane Vect. or Scalar
Temperature range −50/+150◦ Cb −50/+150◦ Cc +20/+200 ◦ C
Typical field range 10 pT–0.1 mTd 100 pT–1 mT 1 fT–1 μTd
Frequency range dc-10 kHz dc-1 kHz dc-1 kHze
Linearity Requires feedback Limited range of H Requires feedback
Size mm–cm 0.1 mm–cm mm–cm
Material Soft ferromagnet Composite Alkali gas
Technology Bulk Thin film or bulk Vapor cells
a SERF-optical magnetometers
b Limited by the soft ferromagnetic coercivity
c Note that some multiferroic materials operate at low temperature (<100 K) [30]
d Can be extended by a closed-loop scheme
e Can be extended by feedback

Noise

Definitions

To fully define a sensor’s performance and its ability to detect a signal defined by
several parameters (field strength, frequency . . . ), the signal-to-noise ratio (SNR) in
the operating conditions is given. Once this value is known, it indicates if the sensor
is adequate for the detection of the targeted signal (if the SNR is larger than 1) or
the number of averages required
√ to extract the signal from the noise (knowing that
the SNR is multiplied by N, N being the number of repeated measurements).
Any macroscopic quantity (such as the output voltage of a sensor) experiences
fluctuations around its mean value, which gives rise to the sensor’s intrinsic noise.
This noise will give the lowest SNR achievable, in absence of external perturbations.
Applying a spectral analysis of the signal (here taking V as an example) via a
Fourier transform allows separating and understanding the various noise sources.
Under Fourier transformation, V can be expressed in the frequency domain as:
1542 M. Pannetier-Lecoeur and C. Fermon

 t0
1
VT = V (t) eiωt dt. (8)
2π 0

The power associated with the fluctuations is:

 t0  
1  
P = V (t)2  dt. (9)
t0 0

From the above, one can express the power spectral density (PSD):
 
VT (ω)2 
SV = lim 2π . (10)
T →∞ t0

It is given in V2 /Hz. The field equivalent noise can be expressed as:


 SV
SB = . (11)
s
√ √
where s is the sensitivity in V/T, and SB is expressed in T/ Hz.

Noise Sources [31]

A precise expression for the PSD and its field equivalent can be only theoretically
given when all the noise sources in the sensor are identified and when their origin is
known, but most of the time, the PSD can only be evaluated from phenomenological
approaches and measurements of the parameters specific to the sensor.
Among the various noise sources, one may cite (for resistor-type sensor) thermal
noise [32] and shot noise (can be fully described theoretically), low-frequency noise
and random telegraph noise; the latter can be only expressed by phenomenological
expressions (Hooge formula [33] for 1/f noise in resistors). For sensors based on
quantum properties, exact expressions can give the theoretical noise limit, such as
in optical magnetometers or SQUIDs.

Detection Schemes

Magnetometer

Magnetic field sensing in magnetometer mode is the simplest; the sensor is used to
record at a single location and its output depends on the H component for a vector
magnetometer, or on the H magnitude for scalar one.
31 Magnetic Sensors 1543

Gradiometer

When the field source is local, it may be very useful to use the sensor in a
gradiometer configuration, which eliminates distant sources and environmental
noise. According to the location of the source and to the sensing direction of the
magnetometer, gradiometers can be axial or radial. When a gradiometer comprises
two magnetometers, one close to the source and the second at a larger distance, it is
a first-order gradiometer. More complex gradiometers (second, third order) integrate
three or four magnetometers to reject higher order of noise contamination and refine
the signal of interest (Figs. 13 and 14).

Fig. 13 Noise spectral density and temporal outputs

Fig. 14 First-order gradiometer: (a) radial along z and (b) axial along x and y
1544 M. Pannetier-Lecoeur and C. Fermon

Fig. 15 Closed-loop measurement: The sensor output is compared to a reference value (fixed
voltage, temperature-compensated voltage, or dummy sensor-sensor insensitive to the external
field). The difference is amplified and sent to a coil through an inverting amplifier. It generates a
field HF which is directly proportional to the applied field H . Appropriate filtering can be inserted
in the feedback loop

Closed Loop

A number of magnetic sensors are inherently nonlinear. This is the case for SQUIDs,
GMI, or optical sensors and to a lesser extent for all magnetoresistive sensors. For
that reason, a closed-loop scheme is often used even in integrated sensors. The prin-
ciple is to have a feedback coil (or wire) which creates a field on the sensor in order
to maintain it in a field where the sensor is the most sensitive. When the external
applied field varies, the current in the feedback coil is adjusted by an analog or
digital electronic loop in order to maintain the sensor output constant. The read value
is not the actual sensor output, it is the current sent to the feedback coil. This scheme
has two advantages; firstly, it linearizes the sensor and secondly, it suppresses sensor
sensitivity variation (as function of temperature, for example) (Fig. 15).

Bridge

Magnetoresistive sensors have a large temperature variation which make them


very difficult to use for dc measurements. For this reason, a half bridge or bridge
configuration, as shown in Fig. 16, are commonly used. These configurations
suppress the main resistance offset and the temperature variation of the main
resistance, but the temperature variation of the sensitivity is not corrected. For that
correction, a closed loop configuration is used.

Applications

Field/Flux Sensing and Flux Transformers

Magnetometers are used for field and flux sensing, either directly when the source
is a magnetic element (bulk magnet, Earth’s field, magnetic nanoparticles . . . ) or
31 Magnetic Sensors 1545

Fig. 16 Bridge configuration (i) with four sensors and (ii) with two sensors and equivalent
resistances RFeed

when the source produces a magnetic field, such as induced by an electrical current
running in a conductive wire or medium.
Field sensors can be used in combination with a flux concentrator, which
transforms the flux caught on a surface to a locally enhanced field in the vicinity of
the sensor. This can be applied by either using soft ferromagnetic flux concentrators
[34, 35] or a superconducting loop combined with a GMR sensor [36], leading to
an increase of sensitivity and/or a change in the direction of the field sensed.

Current Sensing

One of the main commercial applications of magnetic sensors is to detect and/or


measure the current flowing in a wire. This has the strong advantage of measuring
the current without any direct contact to the wire, and without disturbing the
electrical circuit. Magnetic sensors are used on large scale for electricity circuit-
meters, power management, or circuit breakers, but also to detect the tiny currents
carried by neurons and nerves in the body in biomagnetic investigations.

Position Sensing

Magnetic sensors are also widely used, in particular in the automotive market, to
detect the position of a moving element (such as a gear or a crank) either by placing
the magnetic sensor in front of the moving magnetized element or by using a back
bias solution where the magnet is placed on the sensor, and where the moving
or rotating metallic element is changing the magnetization landscape seen by the
sensor. The repetition of a known pattern can be detected and indicates the speed of
the target (Fig. 17).
1546 M. Pannetier-Lecoeur and C. Fermon

Fig. 17 Speed sensor. (a) The sensor is facing a rotating magnetized target. (b) A magnet, placed
on the back of the sensor, generates a field landscape which changes when the metallic target,
containing specific features – here teeth – rotates. (c) Sensor output as a function of the target
rotation angle for (a) and (b) configurations. Plateau or edge detection gives the rotation speed of
the target

Fig. 18 Field range for various sensors. Field sensors are shown in the upper part of the figure
and flux sensors in the bottom part. Note that optical magnetometers usually require heating up to
180 ◦ C and SQUIDs require liquid helium (4 K) or liquid nitrogen (77 K) temperature to operate

Figure 18 summarizes the field range for various sensors based on the physical
principles presented in this chapter. Field sensors (on the top part of the figure) have
in general lower sensitivity than flux sensors, but can be miniaturized.
Table 4 gives an overview of sensitivity and noise for the main sensors type
presented in this chapter, detailing also their maximal measurable field as well as
their main advantages and drawbacks.
31 Magnetic Sensors 1547

Table 4 Comparative table of sensitivity and noise


Intrinsic
Sensor sensitivity Field equivalent noise Maximal Advantages Drawbacks
Standard
 √   √  Best Measurable
(V/T) T/ Hz T/ Hz Field
Hall 1 (Si)- 1 μT 100 nT 10 T Cheap Low sen-
5(InSb) [38, 39] CMOS sitivity
[37] integra-
tiona
AMR 10–50 10 nT– 50 pT 1 mT Robust Linearity
200 pT [41] interme-
[40] diate
sensitiv-
ity
GMR 50–200b 100 pT 20 pT 1–10 mTb Good Saturates
[40] sensitivity at high
fields
CMOS
integration 1/f noise
TMR 1000 50– 20 pT 1–10 mT Very low Saturates
100 pT [44] consump- at high
[42]c tion fields
[43] Good 1/f noise
sensitivity
CMOS
integration
Optical Not 10 fT 0.2 fT 1 μT Very high Integration
pumping relevantd [45] sensitivity dynamic
range
GMI 100 200 pT- 1 pT [46] 100 μT High Linearity
sensitivity
Magnetoelectric 10–1000 1 nT 30 pT 1 mT Passive Process
[47] reliability
Coil Proportional 1 fT at 0.01 fT at 10 T Cheap, No dc
to 1 kHz 1 kHze versatile detection
frequency
and size
Search coil 10,000 100 fT at 10 mT Very No dc
1 Hz [48] sensitive detection
Fluxgate 400– 5 nT 1 pT [6] 100 μT Very Frequency
10,000f sensitive limitation
no 1/f
noise
(continued)
1548 M. Pannetier-Lecoeur and C. Fermon

Table 4 (continued)
Intrinsic
Sensor sensitivity Field equivalent noise Maximal Advantages Drawbacks
Standard
 √  Best  √  Measurable
(V/T) T/ Hz T/ Hz Field
SQUIDS 106 –107 10 fT 1 fT [7] 10 μT Ultrasensitive Cryogenic
cooling
a For Si-Hall sensors
b Sensitivityand range are dependent and can be tuned by shape anisotropy for instance
c Optical detection
d For tuned coils
e Depends on the number of windings
f With flux concentrators

References
1. Ripka, P.: Magnetic Sensors and Magnetometers. Artech House Publishers, Boston (2001).
ISBN: 978-1-58053-057-6
2. Robbes, D.: Highly sensitive magnetometers a review. Sensors Actuators A Phys. 129(1–2),
86–93 (2006)
3. Tumanski, S.: Induction coil sensors a review. Meas. Sci. Technol. 18(3), R31 (2007)
4. Coillot, C., Leroy, P.: Chapter 3: Induction magnetometers principle, modeling and ways of
improvement. In: Magnetic Sensors- Principles and Applications. InTech, London (2012)
5. Butta, M.: Chapter 2: Orthognal fluxgates. In: Magnetic Sensors- Principles and Applications.
InTech, London (2012)
6. http://www.bartington.com, 2015
7. Clarke, J., Braginski, A.I.: The SQUID Handbook: Vol. I Fundamentals and Technology of
SQUIDs and SQUID Systems. Wiley-VCH, Zurich (2004)
8. Clarke, J., Braginski, A.I.: The SQUID Handbook: Vol. II Applications of SQUIDs and
SQUIDs Systems. Wiley-VCH, Zurich (2006)
9. Aine, C.J., Supek, S.: Magnetoencephalography: From Signals to Dynamic Cortical Networks.
Springer (2014)
10. M.R.J. Gibbs and P.T. Squire. Magnetic and magnetoelastic properties of amorphous ribbons
and wires. In Magnetic Ribbons and Wires in Power, Electronic and Automotive Applications,
IEE Colloquium on, pp. 4/1–4/4, Nov 1990
11. Hall, E.H.: On a new action of the magnet on electric currents. Am. J. Math. 2(3), 287–292
(1879)
12. Thomson, W.: On the electro-dynamic qualities of metals:–effects of magnetization on the
electric conductivity of nickel and of iron. Proc. R. Soc. Lond. 8, 546–550 (1856)
13. Coey, J.M.D.: Magnetism and Magnetic Materials. Cambridge University Press (2010)
14. Baibich, M.N., Broto, J.M., Fert, A., Nguyen Van Dau, F., Petroff, F., Etienne, P., Creuzet, G.,
Friederich, A., Chazelas, J.: Giant magnetoresistance of (001)fe/(001)cr magnetic superlattices.
Phys. Rev. Lett. 61, 2472–2475 (1988)
15. Dieny, B., Speriosu, V.S., Parkin, S.S.P., Gurney, B.A., Wilhoit, D.R., Mauri, D.: Giant
magnetoresistive in soft ferromagnetic multilayers. Phys. Rev. B. 43, 1297–1300 (1991)
16. Yuasa, S., Nagahama, T., Fukushima, A., Suzuki, Y., Ando, K.: Giant room-temperature
magnetoresistance in single-crystal Fe/MgO/Fe magnetic tunnel junctions. Nat. Mater. 3(12),
868–871 (2004)
31 Magnetic Sensors 1549

17. Parkin, S.S.P., Kaiser, C., Panchula, A., Rice, P.M., Hughes, B., Samant, M., Yang, S.-H.: Giant
tunnelling magnetoresistance at room temperature with MgO (100) tunnel barriers. Nat. Mater.
3(12), 862–867 (2004)
18. Bloom, A.L.: Principles of operation of the rubidium vapormagnetometer. Appl. Opt. 1(1),
61–68 (1962)
19. Dupont-Roc, J., Haroche, S., Cohen-Tannoudji, C.: Detection of very weak magnetic fields
(10−9 gauss) by 87 Rb zero-field level crossing resonances. Phys. Lett. A. 28(9), 638–639 (1969)
20. Dehmelt, H.G.: Modulation of a light beam by precessing absorbing atoms. Phys. Rev. 105,
1924–1925 (1957)
21. Kominis, I.K., Kornack, T.W., Allred, J.C., Romalis, M.V.: A subfemtotesla multichannel
atomic magnetometer. Nature. 422(6932), 596–599 (2003)
22. Budker, D., Romalis, M.: Optical magnetometry. Nat. Phys. 3(4), 227–234 (2007)
23. Clark Griffith, W., Knappe, S., Kitching, J.: Femtotesla atomic magnetometry in a microfabri-
cated vapor cell. Opt. Express. 18(26), 27167–27172 (2010)
24. Beach, R.S., Berkowitz, A.E.: Giant magnetic field dependent impedance of amorphous fecosib
wire. Appl. Phys. Lett. 64(26), 3652–3654 (1994)
25. Rao, K.V., Humphrey, F.B., Costa Kramer, J.L.: Very large magnetoimpedance in amorphous
soft ferromagnetic wires (invited). J. Appl. Phys. 76(10), 6204–6208 (1994)
26. Hans Hauser, L.K., Ripka, P.: Giant magnetoimpedance sensors. IEEE Instrum. Meas. Mag.
4(2), 28–32 (2001)
27. Phan, M.-H., Peng, H.-X.: Giant magnetoimpedance materials: fundamentals and applications.
Prog. Mater. Sci. 53(2), 323–420 (2008)
28. Vazquez, M., Chiriac, H., Zhukov, A., Panina, L., Uchiyama, T.: On the state-of-the-art in
magnetic microwires and expected trends for scientific and technological studies. Phys. Status
Solidi A Appl. Mater. Sci. 208(3), 493–501 (2011)
29. Zhukov, A., Ipatov, M., Churyukanova, M., Kaloshkin, S., Zhukova, V.: Giant magne-
toimpedance in thin amorphous wires: from manipulation of magnetic field dependence to
industrial applications. J. Alloys Compd. 586, Supplement 1(0):S279–S286 (2014). {SI}:
{ISMANAM} 2012
30. Eerenstein, W., Mathur, N.D., Scott, J.F.: Multiferroic and magnetoelectric materials. Nature.
442(7104), 759–765 (2006)
31. de Freitas, S.C., Mukhopadhyay, S.C., Reig, C. (eds.): Giant Magnetoresistance (GMR)
Sensors, pp. 47–70. Springer, Berlin, Heidelberg (2013)
32. Nyquist, H.: Thermal agitation of electric charge in conductors. Phys. Rev. 32, 110–113 (Jul
1928)
33. Hooge, F.N., Hoppenbrouwers, A.: 1/f noise in continuous gold films. Physica. 45, 386
(1969)
34. Trindade, I.G., Oliveira, J., Fermento, R., Sousa, J.B., Cardoso, S., Freitas, P.P., Raghunathan,
A., Snyder, J.E.: Soft thin films for flux concentrators. IEEE Trans. Magn. 45(1), 168–171
(2009)
35. Clark Griffith, W., Jimenez-Martinez, R., Shah, V., Knappe, S., Kitching, J.: Miniature atomic
magnetometer integrated with flux concentrators. Appl. Phys. Lett. 94(2) (2009)
36. Pannetier, M., Fermon, C., Le Goff, G., Simola, J., Kerr, E.: Femtotesla magnetic field
measurement with magnetoresistive sensors. Science. 304(5677), 1648–1650 (2004)
37. Lenz, J., Edelstein, A.S.: Magnetic sensors and their applications. IEEE Sensors J. 6(3), 631–
649 (2006)
38. Kerlain, A., Mosser, V.: Dynamic low-frequency noise cancellation in quantum well hall
sensors (qwhs). Sensors Actuators A Phys. 142(2), 528–532 (2008) The sixth European
Magnetic Sensor and Actuator conference The Sixth European Magnetic Sensor and Actuator
conference
39. Kerlain, A., Mosser, V.: Low frequency noise suppression in iii–v hall magnetic microsystems
with integrated switches. Sens. Lett. 5(1), 192–195 (2007)
1550 M. Pannetier-Lecoeur and C. Fermon

40. Stutzke, N.A., Russek, S.E., Pappas, D.P., Tondra, M.: Low-frequency noise measurements on
commercial magnetoresistive magnetic field sensors. J. Appl. Phys. 97, 10Q107 (2005)
41. Zimmermann, E., Verweerd, A., Glaas, W., Tillmann, A., Kemna, A.: An AMR sensor-based
measurement system for magnetoelectrical resistivity tomography. IEEE Sensors J. 5(2), 233–
241 (2005)
42. Cardoso, S., Leitao, D.C., Gameiro, L., Cardoso, F., Ferreira, R., Paz, E., Freitas, P.P.: Magnetic
tunnel junction sensors with ptesla sensitivity. Microsyst. Technol. 20(4–5), 793–802 (2014)
43. E. Paz, S. Serrano-Guisan, R. Ferreira, P.P. Freitas.: J. Appl. Phys. 115, 17E501 (2014). https://
doi.org/10.1063/1.4859036
44. Valadeiro, J.P., Amaral, J., Leitao, D.C., Ferreira, R., Cardoso, S.-s.F., Freitas, P.J.P.: Strategies
for pTesla field detection using magnetoresistive sensors with a soft pinned sensing layer. IEEE
Trans. Magn. 51(1), 1 (2015)
45. Dang, H.B., Maloof, A.C., Romalis, M.V.: Ultrahigh sensitivity magnetic field and magnetiza-
tion measurements with an atomic magnetometer. Appl. Phys. Lett. 97(15), 151110 (2010)
46. Portalier, E., Dufay, B., Saez, S., Dolabdjian, C.: Noise behavior of high sensitive GMI-based
magnetometer relative to conditioning parameters. IEEE Trans. Magn. 51(1), 1–4 (2015)
47. Zhuang, X., Sing, M.L.C., Dolabdjian, C., Wang, Y., Finkel, P., Li, J., Viehland, D.: Mechanical
noise limit of a strain-coupled mag- neto(elasto)electric sensor operating under a magnetic or
an electric field modulation. IEEE Sensors J. 15(3), 1575–1587 (2015)
48. Ripka, P., Janosek, M.: Advances in magnetic field sensors. IEEE Sensors J. 10(6), 1108–1116
(2010)
49. Staudacher, T., Shi, F., Pezzagna, S., Meijer, J., Du, J., Meriles, CA., Reinhard, F., Wrachtrup,
J.: Nuclear magnetic resonance spectroscopy on a (5-Nanometer)3 sample volume. Science.
339(6119), 561–563 (2013). https://doi.org/10.1126/science.1231675
50. Tetienne, J.-P., Hingant, T., Martinez, L. J., Rohart, S., Thiaville, A., Diez, L. H., Garcia,
K., Adam, J.-P., Kim, J.-V., Roch, J.-F., Miron, I. M., Gaudin, G., Vila, L., Ocker, B.,
Ravelosona, D., Jacques, V.: The nature of domain walls in ultrathin ferromagnets revealed
by scanning nanomagnetometry. Nature Communications. 6, 6733 (2015). https://doi.org/10.
1038/ncomms7733
51. Rondin, L., Tetienne, J.-P., Hingant, T., Roch, J.-F., Maletinsky, P., Jacques, V.: Magnetometry
with nitrogen-vacancy defects in diamond. Reports on Progress in Physics. 77(5) 056503
(2014). https://doi.org/10.1088/0034-4885/77/5/056503
52. Balasubramanian, G., Chan, I. Y., Kolesov, R., Al-Hmoud, M., Tisler, J., Shin, C., Kim, C.,
Wojcik, A., Hemmer, P. R., Krueger, A., Hanke, T., Leitenstorfer, A., Bratschitsch, R., Jelezko,
F., Wrachtrup, J.: Nanoscale imaging magnetometry with diamond spins under ambient
conditions. Nature. 455, 648 (2008). https://doi.org/10.1038/nature07278
53. Maze, J. R., Stanwix, P. L., Hodges, J. S., Hong, S., Taylor, J. M., Cappellaro, P., Jiang, L.,
Dutt, M. V. G., Togan, E., Zibrov, A. S., Yacoby, A., Walsworth, R. L., Lukin, M. D.: Nanoscale
magnetic sensing with an individual electronic spin in diamond. Nature. 455, 644 (2008)

Myriam Pannetier-Lecoeur received her Ph.D. from University


of Caen in 1999. She worked at the Vrije Universiteit Amsterdam
and joined SPEC-CEA Saclay in 2001. Her research focuses
on spin electronics, noise measurements, and magnetic sensors.
In particular, she is involved in field sensors for biomedical
applications. She has authored more than 60 articles, 3 book
chapters, and filed 20 patents
31 Magnetic Sensors 1551

Claude Fermon interest is focused on nanomagnetism, biomag-


netism, magnetic resonance imaging, and spin electronics. Dr.
Claude Fermon has received a number of awards, among them
the A. Abragam price for the development of neutron reflectivity
and the A. Poirson price for ultrasensitive magnetic sensors. He
has published 140 papers, 1 book, a number of book chapters, and
30 patents
Magnetic Memory and Logic
32
Wei Han

Contents
Introduction: Magnetic Memory and Logic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1554
Magnetic Hard Drive Memory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1554
Hard Drive and Its Road Map . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1555
Giant Magnetoresistance and RKKY Coupling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1556
Tunneling Magnetoresistance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1557
Magnetic Random-Access Memory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1559
Magnetic Random-Access Memory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1560
Spin-Transfer Torque Magnetic Random-Access Memory . . . . . . . . . . . . . . . . . . . . . . . . 1560
Spin-Orbit Torque Magnetic Random-Access Memory . . . . . . . . . . . . . . . . . . . . . . . . . . . 1562
New-Type Magnetic Memories . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1564
Domain Wall Racetrack Memory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1565
Skrymion Racetrack Memory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1567
Antiferromagnetic Memory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1570
2D van der Waals Magnets Memory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1572
Magnetic Logic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1573
Nanomagnet Logic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1574
Magnetic Domain Wall Logic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1576
Magnetic Tunnel Junction Logic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1578
Spin Current Logic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1578
All Spin Logic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1578
Graphene Spin Logic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1581
Magnetoelectric Spin-Orbit Logic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1583
Spin Wave Logic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1584
Conclusion and Outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1586
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1587

W. Han ()
International Center for Quantum Materials, School of Physics, Peking University, Beijing, China
e-mail: weihan@pku.edu.cn

© Springer Nature Switzerland AG 2021 1553


J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_33
1554 W. Han

Introduction: Magnetic Memory and Logic

Since the institution of the International Olympic Committee in 1894, the three
Latin words “Citius, Altius, and Fortius,” which mean “faster, higher, and stronger,”
are the motto of the Olympic Games. Similarly, “faster, denser, and longer” is the
motto that could be adopted by the scientific community behind the development of
the magnetic memory and logic devices for information technology. The storage
density of the magnetic hard drives used as primary memory has seen more
than a thousandfold increase since the first application of giant magnetoresistance
into the read head around 1997. The integration of memory and logic in the
same device, together with progressive miniaturization, is expected to break the
energy and time constraints of the classic Neumann architecture. This chapter
of magnetic memory and logic focuses on the discussion of using magnetic
materials, magnetoresistance, and spin current for information storage and logic
applications. It is organized as follows: The second section discusses the magnetic
hard drive memory and its road map in terms of the storage density, mainly
arising from the physical discoveries of giant magnetoresistance and tunneling
magnetoresistance. The third section presents the magnetic random-access memory
and its development from current-induced magnetic field switching, then spin-
transfer torque switching, to recent spin-orbit torque switching. The fourth section
discusses the new types of magnetic memories, including domain wall racetrack
memory, skyrmion racetrack memory, antiferromagnetic memory, and 2D van der
Waals magnets memory. The fifth section discusses the logic in magnetic memory
devices, such as nanomagnet logic, domain wall logic, and magnetic tunnel junction
logic. The sixth section discusses the spin current logic devices, including all
spin logic, magnetoelectric spin-orbit logic, and spin wave logic. The summary
and future outlook of magnetic memory and logic are discussed in the final
section.

Magnetic Hard Drive Memory

Magnetic hard drives have been the prime repository of digital data for more than
half a century. It is one of the biggest technological and commercial successes of
spintronics and magnetism so far. Just over 10 years ago, the storage capacity of
all the world’s magnetic disk drives surpassed that of analog data storage devices.
Together with complementary metal-oxide-semiconductor (CMOS) technology,
magnetic hard drives have brought consumer the information storage and computing
for our daily lives. The basics of magnetic hard drive memory and the major physical
discoveries that lead to a higher storage density of the magnetic hard drives, such as
giant magnetoresistance and tunneling magnetoresistance, will be discussed in this
section.
32 Magnetic Memory and Logic 1555

Fig. 1 Magnetic hard drives and its road map. (a) An optical image of the magnetic hard drive and
the read-and-write head. (b) The schematic of the magnetic head and recording medium for a hard
drive. B is the minimum length of the magnetic domains. The GMR read sensor is based on giant
magnetoresistance. The inductive “ring-type” head is used for writing information by switching the
local domain’s magnetization. (Reproduced from [1]). (c) The evolution of the hard drive storage
density from 1996 to 2019. (The illustration of the magnetic medium is reproduced from [2])

Hard Drive and Its Road Map

Magnetic hard drives, first created in 1953 by IBM engineers, have provided the
storage devices with high capacity and low cost. The main parts of a magnetic hard
drive include a magnetic disk platter that has a thin magnetic layer on the surface, the
actuator arm that mechanically controls the motion of the read-and-write head, and
the read-and-write head (Fig. 1a). The write head uses conducting current-induced
magnetic field to switch the magnetic domains of the magnetic materials, and the
read head uses magnetic sensor to read the magnetization as the stored information
(Fig. 1b).
The first step to increase the capacity of a hard disk drive was made by using
the anisotropic magnetoresistive device as the magnetic sensor in 1991 by IBM
[1, 3]. Shortly in 1997, the giant magnetoresistance sensor, a more sensitive and
scalable read technique, was used in the read head, which has led to an increase
of the storage density in hard disk drives by more than two orders of magnitude
to ∼300 GB/in2 in 2007 (Fig. 1c). Then, the tunneling magnetoresistive sensor
1556 W. Han

was incorporated into the read head, which further increased the storage density
of hard disk drives. Many great scientists contributed to the fast increase of the
storage density that has led to the transition of the hard disk to consumer electronics.
Among them, Prof. Albert Fert and Prof. Peter Grunberg won the 2007 Nobel
Prize in physics for their discovery of giant magnetoresistance. Prof. Stuart Parkin
was awarded the 2014 Millennium Technology Prize from Technology Academy
Finland, the prominent award for technological innovation, for his discoveries that
have enabled a thousandfold increase in the storage capacity of magnetic disk drives.
In the following, the giant magnetoresistance and tunneling magnetoresistance will
be discussed.

Giant Magnetoresistance and RKKY Coupling

As discussed earlier, the application of giant magnetoresistance into the read heads
in 1997 greatly contributed to the fast increase of the storage density of hard
disk drives. This is probably the biggest commercial success of spintronics so
far that could detect small magnetic fields sensitively at room temperature. The
discovery of the giant magnetoresistance in magnetic multilayer heterostructures
was made by two independent research groups in the late 1980s [1, 4–6]. The
magnetic multilayer heterostructures consist of the ferromagnetic Fe layers that
are separated by nonmagnetic Cr spacing layers (Fig. 2a), which can be grown
by thin film growth techniques, such as molecular beam epitaxy and magnetron
sputtering. The observed giant magnetoresistance ratio in Fe/Cr multilayers was
shown to be highly dependent on the thickness of Fe and Cr layers. The maximum
of the magnetoresistance ratio (MR = (RAP − RP )/RP × 100) is found to be
∼80% measured on the multilayer heterostructures of (Fe (3 nm)/Cr (0.9 nm))60
(Fig. 2b).
The physical origin of giant magnetoresistance is the spin-dependent scattering at
the interfaces of the magnetic multilayers. At zero magnetic field, the magnetization
states of the neighboring Fe layers are antiparallel to each other. In such states, the
electrons of each channel are slowed down every second magnetic layer, leading
to a high-resistance value. On the other hand, when the external magnetic field is
large enough, the magnetizations of the all the Fe layers point to the same direction.
In such parallel magnetization states, the majority spin-polarized electrons can go
easily through all the Fe layers, leading to a small resistance value.
Later, it was found that the giant magnetoresistance highly depended on the
thickness of the nonmagnetic spacing layers [7]. The multilayers consisting of
Fe/Cr, Co/Ru, and Co/Cr were investigated, and an oscillatory behavior of the
giant magnetoresistance was observed. The experimental results measured on the
Fe/Cr multilayer heterostructures are shown in Fig. 2c, where the maximum of
the giant magnetoresistance was observed for the Cr thickness of ∼1 nm with
the antiparallel exchange coupling between the Fe layers. The oscillatory behavior
of the magnetoresistance has been attributed to the oscillatory exchange coupling
(between ferromagnetic and antiferromagnetic) of the Fe magnetizations as a
function of the thickness of the nonmagnetic layers [8]. This mechanism is identified
32 Magnetic Memory and Logic 1557

Fig. 2 Giant magnetoresistance and RKKY coupling. (a) Illustration of the magnetic multilayer
structures consisting of nanometer-thick magnetic Fe and nonmagnetic spacing Cr layers. (b) The
observation of the giant magnetoresistance effect in Fe/Cr/Fe multilayers. HS represents the exter-
nal magnetic field that aligns the Fe magnetizations from antiparallel to parallel configurations. The
largest giant magnetoresistance is ∼80%. (Reproduced from [4]). (c) The giant magnetoresistance
of Fe/Cr multilayers as a function of the thickness of the Cr spacing layer. Inset: illustration of the
RKKY coupling, the oscillatory exchange coupling between ferromagnetic and antiferromagnetic
coupling as a function of the nonmagnetic layer thickness. (Reproduced from [7])

as RKKY coupling, developed by Ruderman, Kittel, Kasuya, and Yosida, which


refers to the coupling of magnetic moments of the ferromagnetic layers through the
conduction electrons in the nonmagnetic layers.

Tunneling Magnetoresistance

Similar to giant magnetoresistance, which arises from the spin-dependent scattering


across a nonmagnetic metal layer, tunneling magnetoresistance can also be observed
1558 W. Han

Fig. 3 (continued)
32 Magnetic Memory and Logic 1559

due to spin-dependent quantum mechanical tunneling for the case of an insulating


spacer. Early observations of the tunneling magnetoresistance were reported by
Julliere, and later by Moodera et al., using a tunnel barrier of amorphous aluminum
oxide [9, 10]. The tunneling resistance is high when the magnetization of the two
ferromagnetic electrodes are parallel and low for the antiparallel magnetization
configurations.
The observation of room temperature giant TMR (up to 220%) in magnetic
tunnel junctions (MTJs) with a crystalline MgO tunnel barrier by the Parkin
and Yuasa groups [11, 12] was a major step forward. Such MTJs consisted of
highly textured (100)-oriented MgO tunnel barriers and two ferromagnetic CoFe
electrodes (Fig. 3a), which were patterned in situ by shadow mask method.
The typical magnetoresistance curves with various annealing temperatures with a
magnetoresistance ratio up to ∼170% are shown in Fig. 3b. The annealing process
in vacuum increases the tunneling magnetoresistance ratio. Such large tunneling
magnetoresistance arises from the spin filtering effect of single crystalline MgO
tunnel barriers, where the symmetry of electron wave functions plays an important
role [13]. Within the MgO barrier, the tunneling current is carried by evanescent
waves of several well-defined symmetries for crystalline MgO, which interact
with the electrons in the metal with Bloch waves of the same symmetry at the
Fermi level. As shown in Fig. 3c in a (001)-oriented MgO barrier between two Fe
ferromagnetic electrodes, the 1 symmetry is well represented at the Fermi level
in the majority spin direction sub-band and not in the minority one. Consequently,
a good connection of the slowly decaying channel 1 with both electrodes can be
obtained only in their parallel magnetic configuration. The next slowest decay rate
is for states with 5 symmetry. However, both majority and minority Fe channels
have a 2 state that decays very rapidly because there are no real 2 bands near the
Fermi energy of MgO.

Magnetic Random-Access Memory

The first proposal of the magnetic random-access memory (MRAM) based on spin
valves [14] ignited a race to achieve a successful prototyping. The long endurance,
high density, and fast writing time have made MRAM an ideal candidate to replace
silicon-based RAMs [15]. The giant TMR in MgO-based magnetic tunnel junctions


Fig. 3 Giant tunneling magnetoresistance at room temperature with a crystalline MgO tunnel bar-
rier. (a) Transmission electron microscopy photograph of the CoFe/MgO/CoFe junction. (b) Large
tunneling magnetoresistance (TMR) of the CoFe/MgO/CoFe magnetic tunnel junctions annealed
at 120 ◦ C, 360 ◦ C, and 380 ◦ C, respectively. The magnetoresistance curves are asymmetric with
respect to the external magnetic field due to an exchange bias from an antiferromagnetic IrMn
pinning layer. (Reproduced from [11]). (c, d) The MgO spin filtering effect that gives rise to giant
tunneling magnetoresistance ratio. Dominant tunneling states are the fully spin-polarized 1 band
Bloch states. (Reproduced from [13])
1560 W. Han

gave a strong boost to MRAM technology [11, 12]. Recently, many semiconductor
companies have announced the large-scale production of MRAM products. In this
section, the basics of the magnetic random-access memory, the two major advances,
and the recent development of spin-orbit torque magnetic random-access memory
will be discussed.

Magnetic Random-Access Memory

Magnetic random-access memory, a nonvolatile storage memory, is based on


magnetic tunnel junctions as the memory cells (Fig. 4a). Depending on the magnetic
configurations of the ferromagnetic electrodes, a high/low magnetoresistance can be
obtained, which represents a binary information of “1” or “0.” Because of the larger
signal read-out of magnetic tunnel junctions, the initial idea of using spin valves,
which are two ferromagnetic layers sandwiching a nonmagnetic metallic layer, was
abandoned in favor of magnetic tunnel junctions. The reading and writing principles
of MRAMs work as follows:

• For reading, the magnetoresistance of the addressed magnetic tunnel junction cell
is measured between the “bit” and “word” lines [1].
• For writing, a magnetic field, generated by current carrying wires, is used as
the oldest method. The current pulses are sent through one line of each array,
and only at the crossing point of these lines, the resulting magnetic field is high
enough to reorient the magnetization of the free layer.

Using current-generated magnetic fields made the downscaling (hence achieving


high device density) very challenging, since the large stray fields would couple to
multiple devices with unavoidably different write parameters. Spin-transfer torque
has been utilized to switch the magnetization orientation of the free layer in a
magnetic tunnel junction cell. This new kind of magnetic memory is called spin-
transfer torque magnetic random-access memory.

Spin-Transfer Torque Magnetic Random-Access Memory

Spin-transfer torque magnetic random-access memory (STT-MRAM) is the result


of many years of important advances in both physics and materials science [15].
Spin-transfer torque provides an efficient method to switch the magnetization. The
advantages over field switching were also recognized by numerous semiconductor
companies that included STT-MRAM in the research portfolio.
The theoretical prediction of spin-transfer torque was first brought by two
physicists John Slonczewski and Luc Berger [16, 17]. The spin-transfer torque
originates from the transfer of spin angular momentum that happens when spin-
polarized electrical currents interact with the magnetic moments of a ferromagnet,
thus reorienting them when the spin current density is large enough. As illustrated
32 Magnetic Memory and Logic 1561

(a) ‘Bit’ lines 1T/1MTJ cell architecture


MRAM
‘Cross point’
architecture

‘Word’ ‘1’
lines Transistor
‘0’
(b) (c) 10
No. of cycles
9

e- e-
Resistance (kΩ)
8
‘0’
7 ‘1’

6 100 ns
1 cycle operation
Transverse
component 5

F1 F2 4
-10 -8 -6 -4 -2 0 2 4 6 8 10
Current (a.u.)
(d) (e)

m Free Layer

mp Hard Layer

Fig. 4 Magnetic random-access memory and two major advances that contributed to the devel-
opment of the magnetic random-access memory, namely, the discovery of spin-transfer torque
and material engineering of CoFeB with perpendicular magnetic anisotropy. (a) Schematic of
magnetic random-access memory device. “1” corresponds to the high-resistance value when
the two magnetic layers are antiparallel, and “0” corresponds to the low-resistance state for the
parallel magnetization of the two magnetic layers. (Reproduced from [1]). (b) Illustration of the
spin-transfer concept introduced by Slonczewski and Berger [16, 17]. A spin-polarized current
is prepared by a first ferromagnetic layer (F1 ) with an obliquely oriented spin polarization with
respect to the magnetization axis of a second ferromagnetic layer (F2 ), thus applying a spin-
transfer torque acting on F2 magnetization. (Reproduced from [1]). (c) The tunneling resistance
versus current hysteresis loop of a magnetic tunnel junction cell with CoFe/NiFe as the free layer
and 1.0 nm MgO as the tunnel barrier. The No. of cycles are 1, 10, 100, 1K, 10K, 100K, 1M,
10M, 100M, and 1G (from top to down), respectively. (Reproduced from [18]). (d) Illustration
of the MgO-based magnetic tunnel junction with perpendicular magnetic anisotropy. (e) Current-
induced magnetization switching of the magnetic tunnel junction (40 nm diameter) with current
pulse durations of 300 μs and 1.0 s. The top CoFeB layer is 1.0 nm and the bottom CoFeB layer is
1.7 nm. The MgO thickness is 0.85 nm. (Reproduced from [19])
1562 W. Han

in Fig. 4b, the outflowing current from electrode F1 is spin-polarized. When such
current enters the right ferromagnetic electrode (F2), the spin polarization of the
electrons is not aligned with the magnetization of the ferromagnetic layer; thus,
the spins precess rapidly around a momentum-dependent internal field of the right
ferromagnetic electrode. As a result, the component of spin polarization transverses
to the magnetization decays, transferring spin angular momentum to the ferromag-
net. A typical demonstration of the spin-transfer torque switching of a magnetic
tunnel junction with 1 nm single crystalline MgO tunnel barrier is shown in Fig. 4c.
The current generated by the voltage bias switches the magnetic tunnel junction
between the parallel state with low resistance and antiparallel state with high
resistance via the spin-transfer torque physical mechanism. Under positive bias, the
positive current leads to the switching from antiparallel to parallel magnetization
configurations, while negative voltage leads to the opposite switching. The excellent
stability is also demonstrated after 109 cycles using 100 ns pulses of successively
positive and negative current pulses. The long endurance of spin-transfer torque
makes magnetic random-access memory standing out compared to flash memories
(typically ∼105 cycles) [15].
Magnetic electrodes with perpendicular magnetic anisotropy (out-of-plane easy
axis) can improve thermal stability in MRAM devices (Fig. 4d). Prof. Ohno’s
group in Tohoku University fabricated the first CoFeB thin films with perpendicular
magnetic anisotropy on single crystalline MgO tunnel barriers in 2010 [19].
Figure 4e shows the magnetoresistance as a function of the current bias measured
on a Ta/CoFeB/MgO/CoFeB/Ta MTJ with a diameter of 40 nm. The interfacial
anisotropy between CoFeB and MgO is the origin of the thin CoFeB magnetic
easy axis switching from in-plane to perpendicular axis as the CoFeB thickness
decreases below ∼1.5 nm. These results demonstrated MTJs with small size (40 nm
diameter), sufficient thermal stability, low switching current (49 μA), and high
tunnel magnetoresistance ratio (120%).

Spin-Orbit Torque Magnetic Random-Access Memory

Spin-orbit torque magnetic random-access memories (SOT-MRAMs) are a further


improvement over STT-MRAMs taking advantage of the high spin torques that can
occur at the interface between a magnetic layer and a heavy metal, where pure
spin currents can be generated because of a large spin-orbit coupling [20–22]. This
approach could solve the major problem of STT-MRAM where large write currents
can damage the MgO tunnel barriers. The principle of the operation is to separate
the read-and-write circuitry (Fig. 5a). The reading is done through the tunneling
barrier, which is the same as that in spin-transfer torque magnetic random-access
memory, while the writing consists of passing a large enough current through the
metallic under-layer. For example, the spin Hall effect in β-Ta thin film generates a
spin current on the surface of the film, which can switch the magnetization of the
bottom CoFeB electrode as the free layer in a magnetic tunnel junction (Fig. 5b)
32 Magnetic Memory and Logic 1563

Fig. 5 SOT-MRAM. (a) Schematic of the three-terminal SOT-MRAM device. The reading circuit
is separated from the writing current loop, thus eliminating the risk of current-induced breaking
of the MgO tunnel barrier. (b) The large spin-orbit torque from β-Ta thin film that switches
the magnetization of the bottom CoFeB electrode. (c) The tunneling magnetoresistance of the
CoFeB/MgO/CoFeB as a function of the DC current in the Ta layer. At the positive DC current
of ∼0.9 mA, a lower magnetoresistance is achieved, indicating the switching of the CoFeB
magnetization. (Reproduced from [21]). (d) The comparison of the spin Hall angle of topological
insulators, Pt and β-Ta, for switching the CoTb magnetization. (Reproduced from [23])

[21]. A positive DC current in β-Ta layer switches the magnetic tunnel junction
from antiparallel to parallel configurations, while a negative current switches it
to the antiparallel state with high resistance (Fig. 5c). Compared to STT-MRAM,
one disadvantage of SOT-MRAM is its three-terminal feature, which makes it quite
challenging for high density stoarge.
The magnetization switching arising from the spin current torque can be
generated via two physical mechanisms, namely, spin Hall effect [24–27] and
Rashba-Edelstein effect [28–30]. Quantum materials are a class of materials
whose quantum properties stem from a complex interplay between factors such
as reduced dimensionality, quantum confinement, quantum coherence, quantum
fluctuations, topology of wavefunctions, relativistic spin-orbit interactions, and
fundamental symmetries. Recent developments have exhibited many unique spin-
dependent properties of quantum materials, which can be used as promising material
1564 W. Han

Fig. 6 Quantum materials for spin and charge conversion and spin-orbit torque for magnetization
switching applications. These materials include the topological insulator with spin-momentum
locked surface states, Rashba-split two-dimensional electron gases (such as LaAlO3 /SrTiO3 ,
Bi/Ag, etc.), 2D materials, superconductors, as well as non-collinear antiferromagnetic IrMn3 with
Berry curvature-induced large effective spin-orbit coupling. (Reproduced from [31])

candidates for efficient spin and charge conversion (Fig. 6) [31]. For example,
the spin-momentum locked surface states on three-dimensional topological insu-
lators provide a potentially efficient route for the spin current generation and
magnetization switching [23, 32]. The effective spin Hall angle and the switching
energy of topological insulators have exhibited better properties than conventional
metals with large spin-orbit coupling (Fig. 5d). Besides topological insulators, the
Rashba interface states (Fig. 6), such as LaAlO3 /SrTiO3 and Bi/Ag, also exhibited
considerably large spin-orbit torque [33–35]. The effective spin-orbit torque of
LaAlO3 /SrTiO3 can be tuned by a gate voltage [34, 35]. Other quantum materials,
such as two-dimensional material WTe2 [36], superconductors [37], and non-
collinear antiferromagnetic IrMn3 [38] (Fig. 6), also exhibited interesting spin and
charge conversion properties.

New-Type Magnetic Memories

For the purpose of high reliability, performance, and capacity memory, scientists
have never stopped searching for a new type of magnetic memories based on novel
physical mechanisms and materials. This section will describe several emerging
magnetic memories, including the domain wall racetrack memory, skyrmion race-
track memory, antiferromagnetic memory, and 2D magnets memory.
32 Magnetic Memory and Logic 1565

Domain Wall Racetrack Memory

Compared to the two-dimensional feature in hard disk drives where the information
is stored in a magnetic film determined by its magnetization, domain wall magnetic
racetrack memory can be developed as a three-dimensional intelligent memory chip
with similar storage density but faster random read/write speed (Fig. 7) [39]. One
key development toward the domain wall racetrack memory is the moving of a train
of several domain walls by nanosecond current pulses via spin-transfer torque [39].
The current in a magnetic material is spin-polarized, thus generating a spin-transfer
torque on the magnetic domains and derives the domain wall motion (Fig. 7a). The
reading of stored information of domain wall racetrack memory can be realized
using magnetoresistive sensing devices based on magnetic tunnel junctions that are
close to the racetrack nanowires (Fig. 7b). The writing of domain walls (Fig. 7c) can
be done using either the Oersted field of currents passing along neighboring metallic
nanowires or spin-transfer torque from adjacent magnetic nanoelements.
As discussed earlier, the domain wall racetrack memory relies on the current-
driven domain wall motion in the magnetic nanowires. The evolution of the domain
wall racetrack memory can be divided into several major versions (Fig. 8a),
including the first one that uses in-plane magnetic anisotropy nanowires, the
second version based on the perpendicular magnetic anisotropic nanowires, the third
version that uses perpendicular magnetic anisotropic nanowires with adjacent heavy
metals to provide spin-orbit torque that drives high-speed domain wall motion,

Fig. 7 Magnetic domain wall racetrack memory. (a) A vertical-configuration racetrack stores the
information in a U-shaped nanowire. Pulses of spin-polarized current move the entire pattern of
domain walls coherently along the length of the wire past read-and-write elements. (b, c) The
illustration of the reading and writing of the data in magnetic domain wall racetrack memory.
(d) Three-dimensional arrays of racetracks for high-density storage applications. (Reproduced
from [39])
1566 W. Han

Fig. 8 Evolution of racetrack memory and the high-speed domain wall motion. (a) The evolution
of racetrack memory from in-plane magnetic anisotropy nanowire, perpendicular magnetic
anisotropic nanowire, and spin-orbit torque-driven domain wall motion to the high-speed domain
wall motion by both spin-orbit and exchange coupling torques. (Reproduced from [2]). (b) High-
speed domain wall motion of ∼750 m/s using synthetic antiferromagnetic nanowires driven
by exchange coupling torque. (Reproduced from [40]). (c) High-speed domain wall motion of
ferrimagnetic CoTb nanowires driven by magnetic field. A high domain wall velocity that is larger
than 1000 m/s is observed around the angular momentum compensation temperature. (Reproduced
from [41])

and the fourth version that uses synthetic antiferromagnetic coupled nanowires
with exchange coupling torques [2]. One of the major benefits from in-plane to
perpendicular magnetic anisotropic nanowire is the reduced width of the domain
wall, allowing for higher bit density. Both of the first and second versions of
racetrack memories rely on the spin-polarized current in the nanowires themselves
to generate the spin-transfer torque that drives the magnetic domain walls to move.
The domain wall motion direction is opposite to the current pulse direction.
Elevated domain wall velocity is required to achieve high-speed reading and
writing. It has been demonstrated that the perpendicular ferromagnetic domain
wall on an adjacent heavy metal can lead to high-speed domain wall motion
with velocities of ∼350 m/s with the pulsed current densities of ∼108 A cm−2
[42, 43]. For this type of racetrack memory, the domain wall motion and the
32 Magnetic Memory and Logic 1567

pulsed charge current have the same directions, which is different from the early
versions (Fig. 8a). Both the spin Hall effect-induced spin-orbit torque and the
interfacial Dzyaloshinskii-Moriya interaction contribute to the high-speed domain
wall motion. The spin current generated in the heavy metal layer diffuses into the
neighboring magnetic layer and causes the moments within each domain walls to
rotate toward the spin direction accumulated at the interface. The Dzyaloshinskii-
Moriya interaction exchange field, which has same chirality of spin Hall effect, then
provides a torque on these rotated moments that causes all the domain walls to move
in the same direction.
Recently, scientists have pushed the current pulse-driven domain wall velocity
to 750 m/s using synthetic antiferromagnetic nanowires (Fig. 8b) [40]. The high
speed is accounted by an exchange coupling torque on the Néel domain wall, which
is directly proportional to the strength of the antiferromagnetic exchange coupling
between the two sub-layers. The closer the magnetization of the racetrack is zero, the
faster the domain wall moves. Besides, the net small magnetization also removes the
dipolar coupling, which will be helpful for closely packed domain walls and high-
capacity memory. Another approach is using ferrimagnetic materials, such as CoTb
[41], which also has net zero magnetization at the angular momentum compensation
temperature. The domain wall velocity is remarkably enhanced up to ∼1500 m/s
driven by an external magnetic field of 100 mT (Fig. 8c).

Skrymion Racetrack Memory

Skyrmions are topologically stable field configurations, named after the physicist
Tony Skyrme [44]. Magnetic skyrmions are chiral spin structures with whirling
configurations in magnetic materials. For example, a typical Néel-type skyrmion
has the magnetization pointing up on the edges and pointing down in the center,
and the magnetization rotates by 2π around an axis perpendicular to the diameter
(Fig. 9a).
Their robustness against external perturbations and small size, together with the
low current density needed to move them along a magnetic wire, make them suitable
for a racetrack-like memory concept. The stability of nano-size skyrmions, the
generation by current pulses, the detection via electrical means, and the high-speed
motion of the skyrmions are all necessary elements for a real-world implementation
of a skyrmion racetrack memory (Fig. 9a). Due to their “topological” spin texture,
magnetic skyrmions are theoretically able to move past the defects that can be
found at the edges of nanowires. Results from micromagnetic simulations of spin-
polarized current-induced motion of individual skyrmion Co stripes are shown in
Fig. 9b to support this claim [45]. The skyrmion moves away from the notch
structure on the nanowire under a larger current pulse and stops at the notch under
a smaller current pulse.
Recently, a lot of breakthroughs have been achieved in terms of both physical
mechanisms and materials engineering. For example, it has been shown that
skyrmions are stable in the multilayers consisting of a ferromagnet and a material
1568 W. Han

Fig. 9 Skyrmion racetrack


memory. (a) Illustration of the
skyrmion racetrack memory.
The current pulse moves the
skyrmions along the track via
spin-transfer torque. (b) The
movement of a single
skyrmion under current
pulses with the presence of a
notch on the track.
(Reproduced from [45])

with large spin-orbit coupling (Fig. 10a). At the interface of magnetic multilayers,
the Dzyaloshinskii-Moriya interaction results a non-collinear chiral spin texture
[30]. Following the technique, the stabilization of nanoscale skyrmions down to
∼50 nm at room temperature in tri-layer structures has been achieved at the Pt/Co/Ir
interfaces, where the interfacial Dzyaloshinskii-Moriya interaction is enhanced due
to the opposite signs for Pt/Co and Co/Ir interfaces [46].
The electrical generation and detection of magnetic skyrmions have also been
experimentally realized in magnetic multilayer heterostructures. The creation of
magnetic skyrmion in the Ta(5 nm)/Co20 Fe60 B20 (1.1 nm)/TaOx (3 nm) tri-layers
has been performed at room temperature via a current pulse in a geometrical
constriction (Fig. 10b) [47]. After applying an electric current, the chiral bubble
domains on the left side of the sample transform into magnetic skyrmions on
the right side next to the geometrical constriction. Under the current pulses, the
inhomogeneous current distribution at the neck gives rise to an inhomogeneous
spin-orbit effective field, which causes the continuous expansion of the stripe
domains into circular magnetic skyrmions. The electrical detection of sub-100 nm
magnetic skyrmions has been demonstrated via the anomalous Hall resistance
measurement at room temperature in a Hall bar structure sample made of multilayer
films (Ta(5 nm)|Pt(10 nm)|[Co(0.8 nm)|Ir(1 nm)|Pt(1 nm)]20|Pt(3 nm)) [48].
Before the current pulses on the nano-structured sample with a track width of
400 nm (Fig. 10c), a larger anomalous Hall resistance (red symbols) is observed.
This larger resistance state corresponds to the magnetically saturated state, which
is consistent with the magnetic force microscopy measurement (red frame in
Fig. 10c). The current injection induces a single skyrmion formed in the area
of the Hall cross (green frame in Fig. 10c), and a subsequent decrease of the
Hall resistivity is observed (green symbols), corresponding to 254 n cm per
skyrmion.
32 Magnetic Memory and Logic 1569

Fig. 10 Skyrmions in magnetic multilayer heterostructures: the generation and detection via
electrical means and the high-speed motion. (a) The interfacial Dzyaloshinskii-Moriya interaction
(D12 ) in magnetic multilayers with larger spin-orbit coupling, resulting in a non-collinear chiral
spin texture of the magnetic moments. (Reproduced from [30]). (b) The electrical injection of
multiple skyrmions transferred from the magnetic domains via a confined geometrical constriction.
Top and bottom panels represent the polar magnetic optical Kerr images before and after the current
pulse. The sample is made of Ta/CoFeB/TaOx tri-layer. (Reproduced from [47]). (c) The electrical
detection of a single skyrmion via anomalous Hall measurement. (Reproduced from [48]). (d) The
high-speed motion of the skyrmions up to 120 m/s driven by spin-orbit torque of heavy metallic Pt
and Ta. (Reproduced from [49])

The high-speed motion of magnetic skyrmions at room temperature has also been
achieved using current-induced spin-transfer torque [49]. A skyrmion velocity up
to 120 m/s is observed on the [Pt(3 nm)/Co(0.9 nm)/Ta(4 nm)]15 sample under a
current density of 5 × 107 A/cm2 at room temperature characterized by magnetic
transmission soft X-ray microscopy (Fig. 10d). The current-driven skyrmion motion
is due to the Slonczewski-like torque arising from the spin Hall effect in Pt
and Ta, which is similar to the physical origin that drives high-speed Néel-type
domain wall motion in ferromagnetic/Pt bilayer structures [42, 43]. These important
progresses of the stable skyrmions at room temperature without magnetic field,
electrical generation, and detection of magnetic skyrmions, as well as the high-speed
motion, might pave the way for the future racetrack memory based on magnetic
skyrmions.
1570 W. Han

Antiferromagnetic Memory

Compared to ferromagnetic memories, whose magnetic moments can sometimes


be accidentally reoriented by external magnetic fields, antiferromagnetic memories
would be robust against charge and magnetic field perturbations [50–52]. Ideally,
ultrafast spin dynamics of antiferromagnet in the THz range implies the possibility
of magnetization switching in picoseconds time scale [50]. The robustness and
ultrafast switching are two advantages of using antiferromagnets for memory
applications.
Toward an antiferromagnetic memory, the efficient writing and reading informa-
tion via electrical means are necessary. The reading of the magnetic moments in
an antiferromagnet could be based on the anisotropic magnetoresistance effect that
was demonstrated about 10 years ago [53]. Recently, the manipulation of the mag-
netization vectors in the antiferromagnet CuMnAs by electrical current pulses via
field-like spin-orbit torque [52] is a major breakthrough toward antiferromagnetic
memories. Figure 11a shows the crystalline structure of the antiferromagnetic CuM-
nAs. Despite the crystal having an inversion symmetry with the center of inversion
at an interstitial position (green sphere in the figure), the two Mn sub-lattices have
broken inversion symmetry. In such inversion symmetry broken systems, field-like
spin-orbit torque could be generated due to the Rashba-Edelstein effect (Fig. 11b),
which implies that the local non-equilibrium spin polarizations of the two Mn sub-
lattices have opposite signs with respect to each other. Hence, the antiferromagnetic
moments tend to align along the spin polarization direction that is perpendicular to
the applied current. In the CuMnAs antiferromagnetic memory device (Fig. 11c),
the writing current (Jwrite ) was applied along the CuMnAs crystal’s [100] and
[010] directions, indicated by the black and red arrows. The field-like torques
due to the charge current manipulate the antiferromagnetic moment, resulting
the Néel vector direction perpendicular to the writing charge current. To probe
the antiferromagnetic Néel vector changed by the current, both X-ray magnetic
linear dichroism photoemission electron microscopy (Fig. 11d) and anisotropic
magnetoresistance (Fig. 11f) measurements have been used. As shown in Fig. 11e,
after applying three current pulses of amplitude 6.1 × 106 A cm−2 and a duration
of 50 ms, the sign of the X-ray magnetic linear dichroism difference indicates a
rotation of the antiferromagnetic moments toward a direction perpendicular to the
current.
Anisotropic magnetoresistance is used as read-out mechanism of the
magnetic moment by electric means. After three pulses of the charge current
(4 × 106 A cm−2 ) along CuMnAs [100] direction, the transverse resistance (R⊥ )
exhibited a higher resistance value (Fig. 11f). On the other hand, it reached a
lower resistance value after pulses of the charge current (4 × 106 A cm−2 ) along
CuMnAs [010] direction. The correlation between the average domain orientation
and the anisotropy magnetoresistance shows the reproducible switching in response
to orthogonally applied current pulses. The stability of the information storage
using antiferromagnetic memory is further testified in the presence of large external
32 Magnetic Memory and Logic 1571

Fig. 11 Antiferromagnetic memory based on the antiferromagnetic material CuMnAs. (a) The
spin configuration of two sub-lattice Mn atoms in the CuMnAs. The full CuMnAs crystal is
centrosymmetric around the interstitial position highlighted by the green ball. (b) Schematic of
the generation of spin current via spin-galvanic effect. The applied current (J) results in a net
in-plane spin polarization (thick red arrow). (c) The optical image and measurement geometry
of the antiferromagnet CuMnAs device. (d) Schematic for the X-ray magnetic linear dichroism
photoemission electron microscopy measurements. X-rays are incident at 16◦ to the sample
surface, with polarization vector in the film plane. (e) Difference between X-ray magnetic linear
dichroism photoemission electron microscopy images taken after applying alternate orthogonal
current pulse trains of 6.1 MA cm−2 . (f) The transverse resistance (R⊥ ) after applying alternating
current pulses along CuMnAs [100] and [010] directions at 273 K. The current density is
4 × 106 A cm−2 . A constant background is subtracted from transverse resistance. (g) Experimental
proof of the functionality of the antiferromagnetic memory in the presence of an external magnetic
field of 12 T. (Reproduced from [52, 54])

magnetic field. Consistent with the antiferromagnetic nature, even a large external
magnetic field of 12 T is not sufficient to overwrite the signal set by the current
pulses at zero magnetic field (Fig. 11g).
Besides the antiferromagnetic CuMnAs films, other non-centrosymmetric anti-
ferromagnetic Mn2 Au can also be switched by Neel spin-orbit torque, which
exhibits the potential to be used in future antiferromagnetic memories [55].
Furthermore, spin-Hall orbit torque from an adjacent heavy metallic Pt layer has
1572 W. Han

been used to switch the magnetic moments of antiferromagnetic NiO thin films
[56], which might expand the antiferromagnetic material candidates toward memory
applications.

2D van der Waals Magnets Memory

The reduction dimensionality of three-dimensional materials down to 2D is possible


to lead new quantum properties. In ferromagnetism, dimensionality plays an
essential role in determining the impact of thermal fluctuations on the critical
temperatures. Recently, 2D van der Waals magnets associated with strong intrinsic
spin fluctuations have attracted considerable attention [57–60]. These exfoliable
materials provide an ideal platform for exploring magnetism in the 2D limit and
are also attractive for the purposes of high-density memory applications since their
magnetization can be efficiently manipulated via spin-transfer torque and electric
field effect.
Several typical 2D van der Waals ferromagnetic materials have been recently
identified, such as insulating CrI3 (Fig. 12a), semiconducting Cr2 Ge2 Te6 (Fig. 12b),
and metallic Fe3 GeTe2 (Fig. 12c) [61–65]. To probe the 2D ferromagnetism,
magnetic optical Kerr effect is used. The Kerr signal shows different response in
reflection or transmission of a dielectric medium depending on the polarization state
of the incoming photon – which is induced by the presence of magnetic order in the
reflecting material. The magnetic optical Kerr angle of single-layer CrI3 exhibits a
hysteresis as function of the perpendicular magnetic field (Fig. 12d), indicating the
perpendicular anisotropic nature of 2D CrI3 .
Of particular interest is the bilayer CrI3 , for which the interlayer exchange
coupling is antiferromagnetic. Thus, at small magnetic field, the magnetic Kerr
response is almost zero. When the magnetic field is larger than its spin-flip field,
the two CrI3 layers have parallel magnetic moments (Fig. 12d). This interesting
property has been used to fabricate a magnetic tunnel junction that consists of two
layers of graphite and one bilayer CrI3 (Fig. 12e). Owing to the antiferromagnetic
coupling of the magnetization in the adjacent layers, CrI3 tunnel barriers exhibit
extremely large magnetoresistance in excess of 10,000%, occurring in a series of
sharp jumps (Fig. 12f) [66, 67]. This result can be understood in the sense that the
bilayer CrI3 acts as a unique magnetic tunnel junction that in the antiferromagnetic
state, each layer acts as a tunnel barrier with partial spin filtering for the opposite
spin-polarized electrons in the successive two CrI3 layers. When the magnetic field
is small, the antiferromagnetic coupling of the two layers results in high-resistance
state due to the opposite spin polarization filtering effect. When the magnetic field is
larger, the spin states of the two CrI3 layers become parallel to each other, leading to
strongly reduced spin-filtering effect for the majority spins, and a lower resistance is
observed. The magnetic tunnel junctions consisting two layers of metallic Fe3 GeTe2
and an insulating BN thin barrier, whose structure is similar to MgO-based magnetic
tunnel junction, have also been reported with magnetoresistance up to 160% [68].
32 Magnetic Memory and Logic 1573

Fig. 12 Emergence of 2D van der Waals magnets for future magnetic memory applications.
(a–c) Illustration of several typical 2D van der Waals magnets, CrI3 , Cr2 Ge2 Te6 , and Fe3 GeTe2 .
(Reproduced from [61–63]). (d) The Kerr rotation of the monolayer and bilayer CrI3 . In bilayer
CrI3 , the interlayer exchange coupling between two CrI3 layers is antiferromagnetic, illustrated
by the blue and red spins. (Reproduced from [61]). (e) The schematic of the bilayer CrI3 -based
magnetic tunnel junction. (f) The tunneling conductance as a function of the external magnetic
field that is perpendicular to the CrI3 plane. (Reproduced from [66])

Besides the extremely large magnetoresistance, the large area growth, electrical
manipulation via electrical field and spin-orbit torque, and read-out of the magneti-
zation in 2D van der Waals magnets have also been investigated intensively in the
last 2 years, i.e., see the recent review article [59]. These important progresses of
2D van der Waals magnets might pave an alternative route to achieve high-density
magnetic memories.

Magnetic Logic

Modern electronics rely on the use of CMOS technology for computing. For
the quick growth demand of emerging computing paradigms such as artificial
intelligence, new computing technologies are needed as a workaround to the von
Neumann bottleneck, decrease switching energy, and improve device interconnec-
tion and provide a complete logic and memory family. Nonvolatile logic-in-memory
architecture, where nonvolatile memory elements are distributed over a logic-circuit
plane, is expected to realize both ultra-low-power and reduced interconnection
1574 W. Han

Table 1 The major logic


gates and their
notations/symbols

delay. For the operation of complete logic devices (Table 1), a complete minimal
set is composed of a basic binary operator like logical AND or logical OR and
the unary operator NOT. Recent progresses of logic devices using nano-magnetic
elements and tunneling junctions have been made toward the logic memory in
a single device. This section will discuss the magnetic logic based on the nano-
magnetic dots, magnetic domain walls, and magnetic tunneling junctions as follows.

Nanomagnet Logic

Regarding nanomagnet for logic applications, an array of magnetic quantum dot


cellular automata has recently been used to perform logic operations [69]. The logic
operation uses the magnetization directions of the single-domain magnetic dots,
which couples to their nearest neighbors via magneto-static interactions. The inter-
action also contributes to the information propagating between the nanomagnetic
dots in the networks.
The magnetic quantum dot cellular automata consists a long chain of circular
dots with a single elongated input dot (Fig. 13a) [69]. The nanomagnetic dots are
made of 10-nm-thick magnetic alloy with very small magnetic uniaxial anisotropy
field. A logic “1” or “0” corresponds to the magnetization of the input dot with a
magnetization pointing to the right or the left, which can be set by a single magnetic
field pulse. The logic operations are performed using an applied oscillating magnetic
field which feeds energy into the system and also serves as a clock (Fig. 13b). Under
a weak oscillating magnetic field of 625 Oe and 30 Hz frequency combined with a
210 Oe bias magnetic field that is applied along the chain of dots, no response of
the magnetic optical Kerr effect is observed when the input dots are set to “0.” In
contrast, when the input dots are set to “1,” all of the networks switch in step with
the applied oscillating field. If one assigns the negative phase of the oscillating field
to a logic “0” and the positive phase to a logic “1,” then the response of the networks
32 Magnetic Memory and Logic 1575

Fig. 13 Nanomagnetic logic


operation using magnetic
quantum cellular automata.
(a) Illustration of the
magnetic quantum cellular
automata measured by
scanning electron
microscopy. The left elliptical
dot is the input bit that could
be either “1” or “0.” (b) The
experimental results of the
magnetic quantum cellular
automata networks at room
temperature. From top to
bottom, the time response of
the magnetic fields, number
of dots switching with input
of “0,” and number of dots
switching with input of “1.”
(Reproduced from [69])

is the logical AND of the input dot with the oscillating field as one input and the
elongated dot as the other input. In principle, magnetic solitons carry information
through the networks without loss and so should be able to mediate a logic switch
without dissipation at room temperature.
The magnetic quantum dot cellular automata arrays have been properly struc-
tured to perform the three-input majority gate NAND and NOR logic operations
[70]. The multiple magnetic quantum dots have single-domain state and couple to
each other through dipole-dipole interactions due to the stray fields close the outside
of the magnets. The dipole-dipole interaction gives rise to either ferromagnetically
ordered state (collinear along the axes) or antiferromagnetically ordered state (side
by side in parallel). For the configuration of a central nanomagnet surrounded by
four other nanomagnets, three of them can be used as inputs driven by additional
driver nanomagnets oriented in the x-direction, along the clock-field, and the last
one of them (to the right of the central magnet) is the output (Fig. 14a). After
applying a horizontal clock-field, switching inside the gate begins at the input
magnets and ends at the output magnet. After applying a right-oriented clock-field
of 500 Oe with 30 s rise and fall times, inputs (100), (101), (110), and (111) are
written by the driver magnets. On the other hand, with left-oriented clock-field,
inputs (000), (001), (010), and (011) are written by the driver magnets. The three-
input majority gate can be viewed as a programmable NAND or NOR gate in terms
of the central magnet (Fig. 14b, c).
1576 W. Han

Fig. 14 The majority logic gates using magnetic quantum cellular automata. (a) The functioning
majority gates characterized by magnetic force microscopy. The location of the magnets is drawn
superimposed on the magnetic force microscopy results. Top panel from A to D: Clock-field
applied horizontally to the right. Bottom panel from E to H: Clock-field applied horizontally to
the left. The black insets show alignment of magnetic dipoles, accounting for antiferromagnetic
and ferromagnetic coupling, and demonstrate majority logic gate functionality. (b) NAND gate
truth table. (c) NOR gate truth table. The logic state of the central nanomagnet is determined
by the logical majority vote of its three-input neighbors, of which the ferromagnetically coupled
neighbors vote directly and the antiferromagnetically coupled neighbor votes inversely to its
magnetic state. (Reproduced from [70])

Magnetic Domain Wall Logic

A magnetic domain wall is the interface region between two oppositely aligned
magnetic domains in ferromagnetic materials. The domain wall is highly mobile
and can propagate in a nanowire by spin-polarized current torque. Recently, the
magnetic domains have been constructed into logic circuits consisting of the logic
elements, such as NOT, AND, signal fan-out, and signal cross-over [71]. All these
elements in the logic architecture can be built using planar magnetic nanowires
and domains (Fig. 15a). A single global magnetic field rotates in the plane of the
device and acts as both the clock and the power supply to operate with major logic
functions.
Figure 15b shows a typical integrated logic device consisting of one NOT gate,
one AND gate, two fan-out junctions, and one cross-over junction, which are made
of various domain structures. The magnetic optical Kerr signal at position I exhibits
32 Magnetic Memory and Logic 1577

Fig. 15 Magnetic domain wall logic. (a) Schematic drawing of the optimized magnetic domain
wall logic elements for the typical logic gates and junctions. (b) The magnetic domain wall network
containing one NOT gate, one AND gate, two fan-out junctions, and one cross-over junction. (c)
The magnetic optical Kerr effect measurements from the positions (I–IV shown in (b)) that describe
the operation of the magnetic logic within a counterclockwise rotating field. Traces II and III are
inferred from trace I and show the magnetization state of the AND gate’s input wires. (Reproduced
from [71])

a switching period of three field cycles, indicating the correction operation of the
NOT gate and sequential fan-out elements (Fig. 15c). For the AND gate, the logic
inputs are the magnetization states of the positions II and III, and the output is
the magnetization state at the position IV. The magnetic optical Kerr signal at the
position IV shows a high value, which corresponds to logic “1,” showing the correct
operation of the AND gate.
1578 W. Han

Magnetic Tunnel Junction Logic

Magnetic tunnel junction logic devices have been constructed based on the tunneling
magnetoresistance effect in magnetic tunnel junctions [72]. For this purpose,
MgO-barrier-based magnetic tunnel junctions are used due to the large tunnel
magnetoresistance ratio at room temperature [11, 12]. The logic value of “1” uses
the low-resistance states when the magnetization of two magnetic layers is parallel
(Fig. 16a). On the other hand, the logic value of “0” uses the high-resistance states
when the magnetization of two magnetic layers is antiparallel. These two states can
be switchable using spin-transfer torque (Fig. 16b).
The ADD and CARRY logic-in-memory circuits using magnetic tunnel junctions
and metal oxide semiconductor transistors have been constructed (Fig. 16c). Since
the inputs are stored using the magnetic tunnel junctions, the logic operation is
nonvolatile, and the supply voltage can be cut off with maintaining stored data
in a standby state. This eliminates the static power dissipation of the logic circuit
with stored inputs. Furthermore, a single pinned magnetic tunnel junction-based
logic device can operate the six logic functions AND, OR, NAND, NOR, XOR, and
XNOR by introducing a current input line passing through MTJ as one of the input
[73]. The current input (“1” for current and “0” without current), the Joule heat
raises the temperature of the magnetic tunnel junction to the pinned layer blocking
temperature, and pinned layer magnetization direction can be reversed under the
magnetic field. The logic is read out using a Wheatstone bridge (Fig. 16d), and
the XOR and OR gate logic can be operated after gate SET steps (Fig. 16e). After
the XOR gate SET, the resistance state of logic-active magnetic tunnel junction
corresponds to the logic “1”; the XOR gate logic is operated. When inputs are 1,
1, and 1, the magnetization directions of the free layer and pinned layer become
parallel. On the other hand, the OR gate SET leads to the antiparallel states of the
pinned layer and free layer.

Spin Current Logic

In this section, we discuss the spin current logic, which uses ferromagnetic
electrodes and spin current to carry the information, communicate, and perform
the logic. For the spin current, the angular momentum can be carried by electrons
conducting metals, semiconductors, and magnons/spin waves in ferromagnetic
materials. The following discusses the several recent important progresses of spin
current logic, including all spin logic, graphene spin logic, magnetoelectric spin-
orbit logic, and spin wave logic.

All Spin Logic

The all spin logic devices use nanomagnets as digital spin capacitors to store
information and spin currents in the spin channel/interconnect to communicate
32 Magnetic Memory and Logic 1579

Fig. 16 Magnetic logic based on magnetic tunnel junction devices. (a) Schematic for low-
resistance (“1”) and high-resistance (“0”) states that correspond to the parallel and antiparallel
configurations of the two magnetic layers across the MgO tunnel barrier, respectively. (b) Current-
induced magnetization switching in a typical magnetic tunnel junction. (c) Nonvolatile full
adder based on logic-in-memory architecture. The magnetic tunneling junctions (B and B ) are
merged into logic-circuit planes (surrounded by dotted lines). ((a–c) Reproduced from [72]). (d)
Schematics of Wheatstone bridge-type spin logic based on magnetic tunnel junction devices. (e)
The experimental results of the XOR and OR logic gate operations. ((d, e) Reproduced from [73])

information between the nanomagnets (Fig. 17a) [74]. The magnetization of


the nanomagnets can be switched to present binary information data, i.e., by
spin-transfer torque from another nanomagnet. The spin channel/interconnect is
extremely important for the logic operations, which transmits information from the
input to the output. The injection of pure spin current can be achieved in the nonlocal
1580

Fig. 17 (continued)
W. Han
32 Magnetic Memory and Logic 1581

geometry spin valves [75]. The spin interconnect could be either a conducting
metal or a semiconductor [76–78]. The efficient spin injection into semiconductors
could be realized using tunneling spin injection via either tunnel or Schottky
barriers [78–80].
The nonlocal pure spin-transfer torque could be used to switch the magnetization
of the output nanomagnet, which has already been experimentally realized in the
metal spin valves [81]. This process could change the output magnetization to
be the same as the input magnetization and thus can also be viewed as a COPY
logic operation (Fig. 17b). Changing the polarity of the voltage bias to the input
transfers the COPY to a NOT logic operation. An example of the logic gate
operation with three nanomagnets as the inputs and one output is shown in Fig. 17c.
The superposition of the spin currents injected from the three inputs flowing in
the channel determines the final magnetization state of the output. Such gate can
function as an AND/OR gate. Flipping the magnetization direction of the middle
fixed input magnet toggles the functionality between OR and AND. With the applied
voltage bias onto all the three inputs, the transferred signals in the channel change
from COPY to NOT of the input, so that the gate A can work as NAND/NOR gate.
Importantly, the operation of gate logic A does not affect gate B that is in the idle
state.

Graphene Spin Logic

For all the spin logic applications, a major challenge is to develop a suitable
spin transport channel with long spin lifetime and long spin propagation length.
Graphene is a single layer of carbon sheet, which has low intrinsic spin-orbit
coupling and has attracted a lot interest for this purpose. The electrical spin injection
and detection in graphene spin valves were first achieved at room temperature in
2007 [82], and the spin injection efficiency was enhanced to ∼30% with a large
nonlocal magnetoresistance of ∼130 ohms using MgO tunnel barriers in 2010 [83].
These room temperature spin-dependent physical properties, including long spin
diffusion length, large spin signal, and long spin lifetime, make graphene one of
the most promising material candidates for spin channel material in spin logic
applications [84, 85].
For graphene spin logic device [86], the building block is a magneto-logic gate
consisting of a graphene sheet contacted by five ferromagnetic electrodes (Fig. 18a),


Fig. 17 All spin logic. (a) Schematic for an all spin logic with built-in memory. The information
is stored in the bistable states of magnets. The corresponding inputs and outputs communicate with
each other via spin current flowing in the spin channel. (b) Schematic of input/output magnets for
a COPY process. The spin current torque switches the magnetization of the output magnet. (c)
The schematic for the bistable gates. For gate A, the state of the output is determined by the spin
currents injected into the channel by the three inputs (“Send” magnets). This gate can function as
a NAND/NOR gate or AND/OR gate. For gate B, it is in the idle state. (Reproduced from [74])
1582 W. Han

Fig. 18 Graphene spin logic. (a) Schematic drawing of graphene-based spin logic gate consisting
of a graphene sheet contacted by five ferromagnetic electrodes. Two electrodes (A and D) define
the input states, two electrodes (B and C) define the operation of the gate, and one electrode (M) is
utilized for read-out; Vdd drives the steady-state current, and IM (t) is the transient current response,
which gives the output; CM is a capacitor; IW is the writing current to manipulate the magnetization
direction of each ferromagnetic electrode; Ir (t) is the current used to perturb the magnetization of
the electrode M. (Reproduced from [84]). (b) Illustration of XOR logic device and measurement
geometry. A, B, and M are ferromagnetic electrodes (Co/MgO). R is a nonmagnetic electrode
(Ti/Au) used as ground point. Iout and Vout are the measured current and voltage, respectively.
Rsen is a variable resistor, and Voffs is an AC voltage source. (c) Experimental demonstration
of XOR operation: Iout as a function of in-plane magnetic field. Black (red) curve indicates the
magnetic field (H) sweeps upward (downward). Vertical arrows indicate the magnetization states
of electrodes A and B. Top left inset: truth table of XOR logic operation. ((b, c) Reproduced
from [87])

working as the inputs (A and D), the gates (B and C), and the output (M). These
magnetic electrodes consist of MgO/Py as the tunneling spin injection source and
two nanomagnets of CoFe that can switch the magnetization of Py via spin-transfer
torque. The read-out from M is highly dependent on the magnetization states of B
and C, which can extract the spin current that is injected from A and D and diffuses
to M. For example, if B and C electrodes are under the state “0,” the majority gate
logic of the inputs from A and D will be OR gate. Different gate logic operations
can be achieved by switching the magnetization of the Py layer via spin-transfer
torque from CoFe. Experimentally, XOR logic application has been demonstrated
on a graphene spin valve at room temperature [87]. Three ferromagnetic tunneling
contacts and one nonmagnetic Ti/Au contact were fabricated on the graphene
(Fig. 18b). The output voltage Vout (AC voltage) was measured using standard low-
frequency lock-in techniques, and output current Iout (=Vout /Rsen , Rsen is a sensing
resistor) was determined using a current detection. The experimental demonstration
of the XOR logic operation on the graphene spin valve is shown in Fig. 18c. During
32 Magnetic Memory and Logic 1583

the sweeping of the magnetic field, the magnetization of the two input electrodes
A and B changes, while the magnetization direction of the M electrode is kept
downward. The outcoming current is a logic operation output of the two information
of “0” and “1,” as defined by the inputs of A and B.

Magnetoelectric Spin-Orbit Logic

In the all spin logic devices, a large current density is usually required since the
magnetization switching requires the enough spin-transfer torque. The inefficiency
of switching a magnet with spin currents can be an intrinsic limitation. In contrast,
the number of electrons required to form an electrostatic energy barrier of 1 eV is
low, allowing for low-energy magnetization reversal.
Recently, a new spin logic, namely, magnetoelectric spin-orbit logic, has been
proposed, which is expected to compete with current complementary metal-oxide-
semiconductor technology in terms of switching energy and logic density [88].
In such a spin logic device, the input charge information is first converted to
magnetism/spin information via magnetoelectric effect [89], and then the spin infor-
mation is transformed to a charge output via spin-to-charge conversion (Fig. 19a),
such as inverse spin Hall and Rashba-Edelstein effects [29, 30, 90]. The proposed
interconnect wire transfers the charge output of the left magnetoelectric spin-orbit
logic to the current that can be used to switch the input of the next one (Fig. 19b).
The operation mechanism via the interconnect is as follows: The first step is the
injection of spin-polarized electrons from a ferromagnetic electrode into the material
with large spin-orbit coupling (Fig. 19c), which converts the spin current into a
charge current (Ic (output)) that flows in the interconnect. Then the second step that
is the charge current/voltage switches the magnetization of the ferromagnet layer
that is coupled to the magnetoelectric layer via exchange coupling and exchange
bias (Fig. 19d). The magnetization switching in the right magnetoelectric spin-orbit
logic results in the change of the Ic (out) of the right magnetoelectric spin-orbit
logic. The charge current in the interconnect also alleviates the spin signal decay in
the spin interconnects in all spin logic devices [74]. The one-stage magnetoelectric
spin-orbit logic with three input bits and one output bit is shown in Fig. 19e. The
state of the output is determined by the interconnects with three inputs, similar to
all spin logic gate (Fig. 17c) [74].
Compared to complementary metal-oxide-semiconductor technology, the mag-
netic spin-orbit logic device was reported to have superior switching energy (by
a factor of 10 to 30), lower switching voltage (by a factor of 5), and enhanced
logic density (by a factor of 5) [91]. The progress of the magnetic spin-orbit
logic gates requires further advances in the fields of the magnetoelectric and
spin-orbit effects, including the magnetoelectric switching stability, writing time,
spin-to-charge conversion efficiency, spin state reading, the interconnects and the
complexity, etc. [91].
1584 W. Han

Fig. 19 Magnetoelectric spin-orbit logic. (a) Transduction of state variables for a castable charge-
input and charge-output logic device via magnetoelectric and spin-orbit effects. (b) Schematic for
magnetoelectric spin-orbit logic interconnect with cascaded logic gates. Arrows on the ferromagnet
(red color) show the directions of the input and output currents of the device. (c) Illustration for
the spin injection from a ferromagnetic electrode and spin to charge conversion in a material with
large spin-orbit coupling. (d) Illustration for magnetoelectric effect that couples the charge voltage
with the magnetization of the FM layer. m is the FM magnetization, and HEB and HEC are the
exchange bias and the exchange coupling from the magnetoelectric material to the ferromagnetic
layer, respectively. (e) Schematic for one-stage magnetoelectric spin-orbit logic for majority gates.
(Reproduced from [88])

Spin Wave Logic

Besides the electrons that mediated spin current, spin wave, the collective excita-
tions of the magnetization perturbations in ferromagnetic materials (Fig. 20a), can
also carry spin current during the spin wave propagation. One advantage of spin
wave logic is the possibility of Joule-heat-free spin information transfer over long
distance [92].
The spin waves can be excited in ferromagnetic materials by many means. One
common way is utilizing the inductive microwave technique via an alternating
Oersted field induced around the microwave antennas. For the spin wave logic,
32 Magnetic Memory and Logic 1585

Fig. 20 Ferromagnetic spin wave logic. (a) Schematic for spin wave in a ferromagnet. (b) The
long-distance propagation of spin wave in a ferromagnetic insulator yttrium-iron-garnet. The spin
wave is excited by the microwave pulse and detected via inverse spin Hall effect of the Pt strip that
is 3 mm away from the input antenna. (Reproduced from [93]). (c) Conceptual design of spin wave
logic gates based on a Mach-Zehnder-type spin wave interferometer. (Reproduced from [94]). (d)
All spin wave chip proposed for XOR logic operation. The spin wave densities sent to the transistor
sources S1 and S2 are controlled by the input magnon signals I1 and I2 applied to the gates G1 and
G2 . (Reproduced from [95])

low Gilbert damping materials are favored, such as ferrimagnetic insulator yttrium-
iron-garnet (∼ 10−5 at room temperature). For example, it has been shown that
the spin waves can propagate over 3 mm at room temperature [93]. The spin
waves are excited by a microwave antenna and detected by charge voltage signal
on the Pt electrode via inverse spin Hall effect (Fig. 20b). Mach-Zehnder spin
wave interferometer equipped with current-controlled phase shifters has been used
to construct logic gates (Fig. 20c) [94]. The currents applied to the two inputs
control the phases of the two spin waves from sources. The logic output is defined
by the interference of microwave currents induced by the spin waves; a small
amplitude level due to the destructive interference corresponds to logic “0,” and
a high amplitude level due to the constructive interference defines logic “1.” The
change in the magnitude of the output pulsed signal corresponds to the XNOR logic
functionality. The drawback of the Mach-Zehnder interferometer logic gates is that
it is impossible to combine two logic gates without additional magnon-to-voltage
converters. To build all spin wave logic, the control of the spin wave via another
spin wave current is necessary. Recently, it has been demonstrated that such control
is possible owing to nonlinear magnon-magnon scattering [95]. An XOR gate could
be built in the form of a spin wave interferometer with two magnon transistors
embedded into its arms without any external antennas (Fig. 20d).
1586 W. Han

Fig. 21 Antiferromagnetic
spin wave logic. (a)
Schematic for the spin wave
modes in an antiferromagnet.
The red and blue arrows
represent the two sub-lattice
spins in an antiferromagnet
unit cell. (Reproduced from
[96]). (b) The universal logic
gate using antiferromagnetic
spin waves with the fixed 90◦
spin wave injection. The
instructions (I, J, and K),
inputs (A and B), and outputs
(U and L) are made of the
synthetic antiferromagnetic
domain racetracks. The blue
and red regimes represent the
opposite magnetic domains.
(Reproduced from [97])

Compared to single spin wave mode in ferromagnetic materials, the spin wave
dynamics of a collinear easy-axis antiferromagnet admits two degenerate modes
with opposite circular polarization (Fig. 21a) [98]. Similar to optical light, these
two modes can be recombined into an equivalently orthogonal but linearly polarized
basis. These two modes, distinguished by their opposite polarization, can give
rise to a novel degree of freedom to encode and process information. Recently,
a universal logic has been proposed for logic operations using the polarized spin
wave and the antiferromagnetic domain walls [97]. The physical mechanism relies
on the magnetic gating effect that can modulate the spin wave dispersions in an
antiferromagnet. The universal logic gate (Fig. 21b) consists of three instruction
tracks, two inputs, and one output that are made of ferromagnetic domain wall
racetracks and antiferromagnetic spin wave conducting channel. The outputs “U”
and “L” are synthetic antiferromagnetic coupled domain racetracks, which have
been experimentally achieved already [40]. The universal magnetic logic gate can
realize all unary and binary logic gates with the fixed 90◦ spin wave polarization
injection, which can be pre-processed into the desired polarization degrees of 32◦ ,
77◦ , and 120◦ for the performing all types of majority logic operations [97].

Conclusion and Outlook

In summary, there have been enormous progresses using magnetic materials for
memory and logic applications in the last 30 years. The biggest success toward
commercial product is the use of the giant magnetoresistance and tunneling magne-
32 Magnetic Memory and Logic 1587

toresistance as magnetic sensors in magnetic hard drive disks, which have increased
the storage density more than thousandfold. Recently, the delivery of spin-transfer
torque magnetic random-access memory into the market could be considered as
the second success of spintronic devices that use magnetic tunnel junctions as the
memory cells. Looking forward in the next decades, the dream toward higher-
performance magnetic memories will continue to simulate the discovery of new
physical mechanisms and materials. The domain wall racetrack memory, skyrmion
racetrack memory, antiferromagnetic memory, and 2D van der Waals magnets
memory might hold the promise for the future magnetic memory applications.
Regarding magnetic logic applications, many logic-built-in-memory devices
have been proposed and realized for logic operations in research laboratories.
Although the success of magnetic and spin logic devices as commercial products
might need much more time and effort compared to the magnetic hard drives, the
quick growth demand of new computing technologies will make the research on
magnetic and spin current logic definitely a quite crowded and highly competitive
field. Looking forward, magnetic and spin logic devices will be quickly developed
and might be used in our daily lives in the near future.

Acknowledgments The author acknowledges the funding support of the National Basic Research
Programs of China (Grants 2015CB921104 and 2019YFA0308401), National Natural Science
Foundation of China (NSFC Grants 11574006 and 11974025), and Strategic Priority Research
Program of the Chinese Academy of Sciences (No. XDB28000000).

References
1. Chappert, C., Fert, A., Nguyen Van Dau, F.: The emergence of spin electronics in data storage.
Nat. Mater. 6, 813 (2007)
2. Parkin, S., Yang, S.-H.: Memory on the racetrack. Nat. Nanotechnol. 10, 195 (2015)
3. Moser, A., Takano, K., Margulies, D.T., Albrecht, M., Sonobe, Y., Ikeda, Y., Sun, S., Fullerton,
E.E.: Magnetic recording: advancing into the future. J. Phys. D Appl. Phys. 35, R157 (2002)
4. Baibich, M.N., Broto, J.M., Fert, A., Van Dau, F.N., Petroff, F., Creuzet, G., Friederich, A.,
Chazelas, J.: Giant magnetoresistance of (001)Fe/(001)Cr magnetic superlattices. Phys. Rev.
Lett. 61, 2472 (1988)
5. Binasch, G., Grünberg, P., Saurenbach, F., Zinn, W.: Enhanced magnetoresistance in layered
magnetic structures with antiferromagnetic interlayer exchange. Phys. Rev. B. 39, 4828 (1989)
6. Fert, A.: Nobel lecture: origin, development, and future of spintronics. Rev. Mod. Phys. 80,
1517 (2008)
7. Parkin, S.S.P., More, N., Roche, K.P.: Oscillations in exchange coupling and magnetoresistance
in metallic superlattice structures: Co/Ru, Co/Cr, and Fe/Cr. Phys. Rev. Lett. 64, 2304 (1990)
8. Bruno, P., Chappert, C.: Oscillatory coupling between ferromagnetic layers separated by a
nonmagnetic metal spacer. Phys. Rev. Lett. 67, 1602 (1991)
9. Julliere, M.: Tunneling between ferromagnetic films. Phys. Lett. A. 54, 225 (1975)
10. Moodera, J.S., Kinder, L.R., Wong, T.M., Meservey, R.: Large magnetoresistance at room
temperature in ferromagnetic thin film tunnel junction. Phys. Rev. Lett. 74, 3273 (1995)
11. Parkin, S.S.P., Kaiser, C., Panchula, A., Rice, P.M., Hughes, B., Samant, M., Yang, S.-H.: Giant
tunnelling magnetoresistance at room temperature with MgO (100) tunnel barriers. Nat. Mater.
3, 862 (2004)
1588 W. Han

12. Yuasa, S., Nagahama, T., Fukushima, A., Suzuki, Y., Ando, K.: Giant room-temperature
magnetoresistance in single-crystal Fe/MgO/Fe magnetic tunnel junctions. Nat. Mater. 3, 868
(2004)
13. Butler, W.H., Zhang, X.G., Schulthess, T.C., MacLaren, J.M.: Spin-dependent tunneling
conductance of Fe|MgO|Fe sandwiches. Phys. Rev. B. 63, 054416 (2001)
14. Parkin, S.S.P., Roche, K.P., Samant, M.G., Rice, P.M., Beyers, R.B., Scheuerlein, R.E.,
O’Sullivan, E.J., Brown, S.L., Bucchigano, J., Abraham, D.W., Lu, Y., Rooks, M., Trouilloud,
P.L., Wanner, R.A., Gallagher, W.J.: Exchange-biased magnetic tunnel junctions and applica-
tion to nonvolatile magnetic random access memory (invited). J. Appl. Phys. 85, 5828 (1999)
15. Kent, A.D., Worledge, D.C.: A new spin on magnetic memories. Nat. Nanotechnol. 10, 187
(2015)
16. Slonczewski, J.C.: Current-driven excitation of magnetic multilayers. J. Magn. Magn. Mater.
159, L1 (1996)
17. Berger, L.: Emission of spin waves by a magnetic multilayer traversed by a current. Phys. Rev.
B. 54, 9353 (1996)
18. Kawahara, T., Takemura, R., Miura, K., Hayakawa, J., Ikeda, S., Lee, Y.M., Sasaki, R., Goto,
Y., Ito, K., Meguro, T., Matsukura, F., Takahashi, H., Matsuoka, H., Ohno, H.: 2 Mb SPRAM
(SPin-transfer torque RAM) with bit-by-bit bi-directional current write and parallelizing-
direction current read. IEEE J. Solid State Circuits. 43, 109 (2008)
19. Ikeda, S., Miura, K., Yamamoto, H., Mizunuma, K., Gan, H.D., Endo, M., Kanai, S.,
Hayakawa, J., Matsukura, F., Ohno, H.: A perpendicular-anisotropy CoFeB–MgO magnetic
tunnel junction. Nat. Mater. 9, 721 (2010)
20. Miron, I.M., Garello, K., Gaudin, G., Zermatten, P.-J., Costache, M.V., Auffret, S., Bandiera,
S., Rodmacq, B., Schuhl, A., Gambardella, P.: Perpendicular switching of a single ferromag-
netic layer induced by in-plane current injection. Nature. 476, 189 (2011)
21. Liu, L., Pai, C.-F., Li, Y., Tseng, H.W., Ralph, D.C., Buhrman, R.A.: Spin-torque switching
with the giant spin Hall effect of tantalum. Science. 336, 555 (2012)
22. Brataas, A., Hals, K.M.D.: Spin-orbit torques in action. Nat. Nanotechnol. 9, 86 (2014)
23. Han, J., Richardella, A., Siddiqui, S.A., Finley, J., Samarth, N., Liu, L.: Room-temperature
spin-orbit torque switching induced by a topological insulator. Phys. Rev. Lett. 119, 077702
(2017)
24. Dyakonov, M.I., Perel, V.I.: Spin relaxation of conduction electrons in noncentrosymmetric
semiconductors. Sov. Phys. Solid State. 13, 3023 (1972)
25. Zhang, S.: Spin Hall effect in the presence of spin diffusion. Phys. Rev. Lett. 85, 393 (2000)
26. Hirsch, J.E.: Spin Hall effect. Phys. Rev. Lett. 83, 1834 (1999)
27. Sinova, J., Culcer, D., Niu, Q., Sinitsyn, N.A., Jungwirth, T., MacDonald, A.H.: Universal
intrinsic spin-Hall effect. Phys. Rev. Lett. 92, 126603 (2004)
28. Edelstein, V.M.: Spin polarization of conduction electrons induced by electric current in two-
dimensional asymmetric electron systems. Solid State Commun. 73, 233 (1990)
29. Manchon, A., Koo, H.C., Nitta, J., Frolov, S.M., Duine, R.A.: New perspectives for Rashba
spin-orbit coupling. Nat. Mater. 14, 871 (2015)
30. Soumyanarayanan, A., Reyren, N., Fert, A., Panagopoulos, C.: Emergent phenomena induced
by spin–orbit coupling at surfaces and interfaces. Nature. 539, 509 (2016)
31. Han, W., Otani, Y., Maekawa, S.: Quantum materials for spin and charge conversion. npj
Quantum Mater. 3, 27 (2018)
32. Mellnik, A.R., Lee, J.S., Richardella, A., Grab, J.L., Mintun, P.J., Fischer, M.H., Vaezi, A.,
Manchon, A., Kim, E.A., Samarth, N., Ralph, D.C.: Spin-transfer torque generated by a
topological insulator. Nature. 511, 449 (2014)
33. Sánchez, J.C.R., Vila, L., Desfonds, G., Gambarelli, S., Attané, J.P., De Teresa, J.M., Magén,
C., Fert, A.: Spin-to-charge conversion using Rashba coupling at the interface between non-
magnetic materials. Nat. Commun. 4, 3944 (2013)
34. Lesne, E., Fu, Y., Oyarzun, S., Rojas-Sanchez, J.C., Vaz, D.C., Naganuma, H., Sicoli, G.,
Attane, J.P., Jamet, M., Jacquet, E., George, J.M., Barthelemy, A., Jaffres, H., Fert, A., Bibes,
M., Vila, L.: Highly efficient and tunable spin-to-charge conversion through Rashba coupling
at oxide interfaces. Nat. Mater. 15, 1261 (2016)
32 Magnetic Memory and Logic 1589

35. Song, Q., Zhang, H., Su, T., Yuan, W., Chen, Y., Xing, W., Shi, J., Sun, J., Han, W.: Observation
of inverse Edelstein effect in Rashba-split 2DEG between SrTiO3 and LaAlO3 at room
temperature. Sci. Adv. 3, e1602312 (2017)
36. MacNeill, D., Stiehl, G.M., Guimaraes, M.H.D., Buhrman, R.A., Park, J., Ralph, D.C.: Control
of spin-orbit torques through crystal symmetry in WTe2 /ferromagnet bilayers. Nat. Phys. 13,
300 (2016)
37. Wakamura, T., Akaike, H., Omori, Y., Niimi, Y., Takahashi, S., Fujimaki, A., Maekawa, S.,
Otani, Y.: Quasiparticle-mediated spin Hall effect in a superconductor. Nat. Mater. 14, 675
(2015)
38. Zhang, W., Han, W., Yang, S.-H., Sun, Y., Zhang, Y., Yan, B., Parkin, S.S.P.: Giant facet-
dependent spin-orbit torque and spin Hall conductivity in the triangular antiferromagnet IrMn3 .
Sci. Adv. 2, e1600759 (2016)
39. Parkin, S.S.P., Hyayshi, M., Thomas, L.: Magnetic domain-wall racetrack memory. Science.
320, 190 (2008)
40. Yang, S.-H., Ryu, K.-S., Parkin, S.: Domain-wall velocities of up to 750 m s−1 driven
by exchange-coupling torque in synthetic antiferromagnets. Nat. Nanotechnol. 10, 221
(2015)
41. Kim, K.-J., Kim, S.K., Hirata, Y., Oh, S.-H., Tono, T., Kim, D.-H., Okuno, T., Ham, W.S.,
Kim, S., Go, G., Tserkovnyak, Y., Tsukamoto, A., Moriyama, T., Lee, K.-J., Ono, T.: Fast
domain wall motion in the vicinity of the angular momentum compensation temperature of
ferrimagnets. Nat. Mater. 16, 1187 (2017)
42. Ryu, K.-S., Thomas, L., Yang, S.-H., Parkin, S.S.P.: Chiral spin torque at magnetic domain
walls. Nat. Nanotechnol. 8, 527 (2013)
43. Emori, S., Bauer, U., Ahn, S.-M., Martinez, E., Beach, G.S.D.: Current-driven dynamics of
chiral ferromagnetic domain walls. Nat. Mater. 12, 611 (2013)
44. Skyrme, T.H.R.: A unified field theory of mesons and baryons. Nucl. Phys. 31, 556 (1962)
45. Fert, A., Cros, V., Sampaio, J.: Skyrmions on the track. Nat. Nanotechnol. 8, 152 (2013)
46. Moreau-Luchaire, C., Moutafis, C., Reyren, N., Sampaio, J., Vaz, C.A.F., Van Horne, N.,
Bouzehouane, K., Garcia, K., Deranlot, C., Warnicke, P., Wohlhüter, P., George, J.M., Weigand,
M., Raabe, J., Cros, V., Fert, A.: Additive interfacial chiral interaction in multilayers for
stabilization of small individual skyrmions at room temperature. Nat. Nanotechnol. 11, 444
(2016)
47. Jiang, W., Upadhyaya, P., Zhang, W., Yu, G., Jungfleisch, M.B., Fradin, F.Y., Pearson, J.E.,
Tserkovnyak, Y., Wang, K.L., Heinonen, O., te Velthuis, S.G.E., Hoffmann, A.: Blowing
magnetic skyrmion bubbles. Science. 349, 283 (2015)
48. Maccariello, D., Legrand, W., Reyren, N., Garcia, K., Bouzehouane, K., Collin, S., Cros, V.,
Fert, A.: Electrical detection of single magnetic skyrmions in metallic multilayers at room
temperature. Nat. Nanotechnol. 13, 233 (2018)
49. Woo, S., Litzius, K., Kruger, B., Im, M.-Y., Caretta, L., Richter, K., Mann, M., Krone, A.,
Reeve, R.M., Weigand, M., Agrawal, P., Lemesh, I., Mawass, M.-A., Fischer, P., Klaui, M.,
Beach, G.S.D.: Observation of room-temperature magnetic skyrmions and their current-driven
dynamics in ultrathin metallic ferromagnets. Nat. Mater. 15, 501 (2016)
50. Baltz, V., Manchon, A., Tsoi, M., Moriyama, T., Ono, T., Tserkovnyak, Y.: Antiferromagnetic
spintronics. Rev. Mod. Phys. 90, 015005 (2018)
51. Železný, J., Wadley, P., Olejník, K., Hoffmann, A., Ohno, H.: Spin transport and spin torque in
antiferromagnetic devices. Nat. Phys. 14, 220 (2018)
52. Wadley, P., Howells, B., Železný, J., Andrews, C., Hills, V., Campion, R.P., Novák, V., Olejník,
K., Maccherozzi, F., Dhesi, S.S., Martin, S.Y., Wagner, T., Wunderlich, J., Freimuth, F.,
Mokrousov, Y., Kuneš, J., Chauhan, J.S., Grzybowski, M.J., Rushforth, A.W., Edmonds, K.W.,
Gallagher, B.L., Jungwirth, T.: Electrical switching of an antiferromagnet. Science. 351, 587
(2016)
53. Marti, X., Fina, I., Frontera, C., Liu, J., Wadley, P., He, Q., Paull, R.J., Clarkson, J.D.,
Kudrnovský, J., Turek, I., Kuneš, J., Yi, D., Chu, J.H., Nelson, C.T., You, L., Arenholz, E.,
Salahuddin, S., Fontcuberta, J., Jungwirth, T., Ramesh, R.: Room-temperature antiferromag-
netic memory resistor. Nat. Mater. 13, 367 (2014)
1590 W. Han

54. Grzybowski, M.J., Wadley, P., Edmonds, K.W., Beardsley, R., Hills, V., Campion, R.P.,
Gallagher, B.L., Chauhan, J.S., Novak, V., Jungwirth, T., Maccherozzi, F., Dhesi, S.S.: Imaging
current-induced switching of antiferromagnetic domains in CuMnAs. Phys. Rev. Lett. 118,
057701 (2017)
55. Bodnar, S.Y., Šmejkal, L., Turek, I., Jungwirth, T., Gomonay, O., Sinova, J., Sapozhnik, A.A.,
Elmers, H.J., Kläui, M., Jourdan, M.: Writing and reading antiferromagnetic Mn2 Au by Néel
spin-orbit torques and large anisotropic magnetoresistance. Nat. Commun. 9, 348 (2018)
56. Chen, X.Z., Zarzuela, R., Zhang, J., Song, C., Zhou, X.F., Shi, G.Y., Li, F., Zhou, H.A.,
Jiang, W.J., Pan, F., Tserkovnyak, Y.: Antidamping-torque-induced switching in biaxial
antiferromagnetic insulators. Phys. Rev. Lett. 120, 207204 (2018)
57. Gong, C., Zhang, X.: Two-dimensional magnetic crystals and emergent heterostructure
devices. Science. 363, eaav4450 (2019)
58. Burch, K.S., Mandrus, D., Park, J.-G.: Magnetism in two-dimensional van der Waals materials.
Nature. 563, 47 (2018)
59. Mak, K.F., Shan, J., Ralph, D.C.: Probing and controlling magnetic states in 2D layered
magnetic materials. Nat. Rev. Phys. 1, 646 (2019)
60. Gibertini, M., Koperski, M., Morpurgo, A.F., Novoselov, K.S.: Magnetic 2D materials and
heterostructures. Nat. Nanotechnol. 14, 408 (2019)
61. Huang, B., Clark, G., Navarro-Moratalla, E., Klein, D.R., Cheng, R., Seyler, K.L., Zhong,
D., Schmidgall, E., McGuire, M.A., Cobden, D.H., Yao, W., Xiao, D., Jarillo-Herrero, P., Xu,
X.: Layer-dependent ferromagnetism in a van der Waals crystal down to the monolayer limit.
Nature. 546, 270 (2017)
62. Gong, C., Li, L., Li, Z., Ji, H., Stern, A., Xia, Y., Cao, T., Bao, W., Wang, C., Wang, Y., Qiu,
Z.Q., Cava, R.J., Louie, S.G., Xia, J., Zhang, X.: Discovery of intrinsic ferromagnetism in
two-dimensional van der Waals crystals. Nature. 546, 265 (2017)
63. Fei, Z., Huang, B., Malinowski, P., Wang, W., Song, T., Sanchez, J., Yao, W., Xiao, D.,
Zhu, X., May, A.F., Wu, W., Cobden, D.H., Chu, J.-H., Xu, X.: Two-dimensional itinerant
ferromagnetism in atomically thin Fe3 GeTe2 . Nat. Mater. 17, 778 (2018)
64. Xing, W., Chen, Y., Odenthal, P.M., Zhang, X., Yuan, W., Su, T., ong, Q., Wang, T., Zhong, J.,
Jia, S., Xie, X.C., Li, Y., Han, W.: Electric field effect in multilayer Cr2 Ge2 Te6 : a ferromagnetic
2D material. 2D Mater. 4, 024009 (2017)
65. Deng, Y., Yu, Y., Song, Y., Zhang, J., Wang, N.Z., Wu, Y.Z., Zhu, J., Wang, J., Chen, X.H.,
Zhang, Y.: Gate-tunable room-temperature ferromagnetism in two-dimensional Fe3 GeTe2 .
Nature. 563, 94 (2018)
66. Klein, D.R., MacNeill, D., Lado, J.L., Soriano, D., Navarro-Moratalla, E., Watanabe, K.,
Taniguchi, T., Manni, S., Canfield, P., Fernández-Rossier, J., Jarillo-Herrero, P.: Probing
magnetism in 2D van der Waals crystalline insulators via electron tunneling. Science. 360,
1218 (2018)
67. Song, T., Cai, X., Tu, M.W.-Y., Zhang, X., Huang, B., Wilson, N.P., Seyler, K.L., Zhu, L.,
Taniguchi, T., Watanabe, K., McGuire, M.A., Cobden, D.H., Xiao, D., Yao, W., Xu, X.: Giant
tunneling magnetoresistance in spin-filter van der Waals heterostructures. Science. 360, 1214
(2018)
68. Wang, Z., Sapkota, D., Taniguchi, T., Watanabe, K., Mandrus, D., Morpurgo, A.F.: Tunneling
spin valves based on Fe3 GeTe2 /hBN/Fe3 GeTe2 van der Waals heterostructures. Nano Lett. 18,
4303 (2018)
69. Cowburn, R.P., Welland, M.E.: Room temperature magnetic quantum cellular automata.
Science. 287, 1466 (2000)
70. Imre, A., Csaba, G., Ji, L., Orlov, A., Bernstein, G.H., Porod, W.: Majority logic gate for
magnetic quantum-dot cellular automata. Science. 311, 205 (2006)
71. Allwood, D.A., Xiong, G., Faulkner, C.C., Atkinson, D., Petit, D., Cowburn, R.P.: Magnetic
domain-wall logic. Science. 309, 1688 (2005)
72. Matsunaga, S., Hayakawa, J., Ikeda, S., Miura, K., Hasegawa, H., Endoh, T., Ohno, H.,
Hanyu, T.: Fabrication of a nonvolatile full adder based on logic-in-memory architecture using
magnetic tunnel junctions. Appl. Phys. Express. 1, 091301 (2008)
32 Magnetic Memory and Logic 1591

73. Wang, J., Meng, H., Wang, J.-P.: Programmable spintronics logic device based on a magnetic
tunnel junction element. J. Appl. Phys. 97, 10D509 (2005)
74. Behin-Aein, B., Datta, D., Salahuddin, S., Datta, S.: Proposal for an all-spin logic device with
built-in memory. Nat. Nanotechnol. 5, 266 (2010)
75. Johnson, M., Silsbee, R.H.: Interfacial charge-spin coupling: injection and detection of spin
magnetization in metals. Phys. Rev. Lett. 55, 1790 (1985)
76. Jedema, F.J., Filip, A.T., van Wees, B.J.: Electrical spin injection and accumulation at room
temperature in an all-metal mesoscopic spin valve. Nature. 410, 345 (2001)
77. Kimura, T., Otani, Y.: Large spin accumulation in a permalloy-silver lateral spin valve. Phys.
Rev. Lett. 99, 196604 (2007)
78. Lou, X., Adelmann, C., Crooker, S.A., Garlid, E.S., Zhang, J., Reddy, K.S.M., Flexner, S.D.,
Palmstrom, C.J., Crowell, P.A.: Electrical detection of spin transport in lateral ferromagnet-
semiconductor devices. Nat. Phys. 3, 197 (2007)
79. Appelbaum, I., Huang, B., Monsma, D.J.: Electronic measurement and control of spin transport
in silicon. Nature. 447, 295 (2007)
80. Zhou, Y., Han, W., Chang, L.-T., Xiu, F., Wang, M., Oehme, M., Fischer, I.A., Schulze, J.,
Kawakami, R.K., Wang, K.L.: Electrical spin injection and transport in germanium. Phys. Rev.
B. 84, 125323 (2011)
81. Yang, T., Kimura, T., Otani, Y.: Giant spin-accumulation signal and pure spin-current-induced
reversible magnetization switching. Nat. Phys. 4, 851 (2008)
82. Tombros, N., Jozsa, C., Popinciuc, M., Jonkman, H.T., van Wees, B.J.: Electronic spin transport
and spin precession in single graphene layers at room temperature. Nature. 448, 571 (2007)
83. Han, W., Pi, K., McCreary, K.M., Li, Y., Wong, J.J.I., Swartz, A.G., Kawakami, R.K.:
Tunneling spin injection into single layer graphene. Phys. Rev. Lett. 105, 167202 (2010)
84. Han, W., Kawakami, R.K., Gmitra, M., Fabian, J.: Graphene spintronics. Nat. Nanotechnol. 9,
794 (2014)
85. Roche, S., Åkerman, J., Beschoten, B., Charlier, J.-C., Chshiev, M., Dash, S.P., Dlubak, B.,
Fabian, J., Fert, A., Guimarães, M., Guinea, F., Grigorieva, I., Schönenberger, C., Seneor, P.,
Stampfer, C., Valenzuela, S.O., Waintal, X., van Wees, B.: Graphene spintronics: the European
Flagship perspective. 2D Mater. 2, 030202 (2015)
86. Dery, H., Wu, H., Ciftcioglu, B., Huang, M., Song, Y., Kawakami, R., Shi, J., Krivorotov,
I., Zutic, I., Sham, L.J.: Nanospintronics based on magnetologic gates. IEEE Trans. Electron
Devices. 59, 259 (2012)
87. Wen, H., Dery, H., Amamou, W., Zhu, T., Lin, Z., Shi, J., Žutić, I., Krivorotov, I., Sham, L.J.,
Kawakami, R.K.: Experimental demonstration of XOR operation in graphene magnetologic
gates at room temperature. Phys. Rev. Appl. 5, 044003 (2016)
88. Manipatruni, S., Nikonov, D.E., Lin, C.-C., Gosavi, T.A., Liu, H., Prasad, B., Huang, Y.-L.,
Bonturim, E., Ramesh, R., Young, I.A.: Scalable energy-efficient magnetoelectric spin–orbit
logic. Nature. 565, 35 (2019)
89. Spaldin, N.A., Fiebig, M.: The renaissance of magnetoelectric multiferroics. Science. 309, 391
(2005)
90. Sinova, J., Valenzuela, S.O., Wunderlich, J., Back, C.H., Jungwirth, T.: Spin Hall effects. Rev.
Mod. Phys. 87, 1213 (2015)
91. Manipatruni, S., Nikonov, D.E., Young, I.A.: Beyond CMOS computing with spin and
polarization. Nat. Phys. 14, 338 (2018)
92. Chumak, A.V., Vasyuchka, V.I., Serga, A.A., Hillebrands, B.: Magnon spintronics. Nat. Phys.
11, 453 (2015)
93. Chumak, A.V., Serga, A.A., Jungfleisch, M.B., Neb, R., Bozhko, D.A., Tiberkevich, V.S.,
Hillebrands, B.: Direct detection of magnon spin transport by the inverse spin Hall effect.
Appl. Phys. Lett. 100, 082405 (2012)
94. Schneider, T., Serga, A.A., Leven, B., Hillebrands, B., Stamps, R.L., Kostylev, M.P.: Realiza-
tion of spin-wave logic gates. Appl. Phys. Lett. 92, 022505 (2008)
95. Chumak, A.V., Serga, A.A., Hillebrands, B.: Magnon transistor for all-magnon data processing.
Nat. Commun. 5, 4700 (2014)
1592 W. Han

96. Cheng, R., Daniels, M.W., Zhu, J.-G., Xiao, D.: Antiferromagnetic spin wave field-effect
transistor. Sci. Rep. 6, 24223 (2016)
97. Yu, W., Lan, J., Xiao, J.: Magnetic logic gate based on polarized spin waves. Phys. Rev. Appl.
13, 024055 (2020)
98. Keffer, F., Kittel, C.: Theory of antiferromagnetic resonance. Phys. Rev. 85, 329 (1952)

Wei Han received his PhD from the University of California,


Riverside, in 2012. He worked at the IBM Almaden Research
Center from 2012 to 2014. In 2014, he joined the Interna-
tional Center for Quantum Materials at Peking University. He
is currently a tenured associate professor, and his research lab
focuses on spintronics in emergent materials for spintronic device
applications.
Magnetochemistry and Magnetic
Separation 33
Peter Dunne

Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1594
Magnetic Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1594
Magnetic Susceptibility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1595
Magnetic Susceptibility of Liquids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1598
Pascal’s Constants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1598
Magnetic Field Effects in Chemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1599
Magnetoelectrochemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1601
Electrochemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1602
Magnetic Forces in a Uniform Magnetic Field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1605
Magnetic Forces in a Non-uniform Magnetic Field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1610
Magnetic Water Treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1614
Magnetic Separation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1616
High-Gradient Magnetic Separation (HGMS) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1618
Magnetic Carriers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1618
Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1621
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1623

Abstract

The effects of magnetic fields in chemistry and electrochemistry, along with


methods of magnetic separation, are covered here. First there will be a gener-
alized introduction to relevant aspects of magnetism, magnetic susceptibilities,
and magnetochemistry. Then follows a short account of the fundamentals of elec-
trochemistry, followed by a development of the various magnetic effects. In the
second part, magnetic separation in industry for both refinement and recycling

P. Dunne ()
Institut de Physique et de Chimie des Matériaux de Stasbourg, Strasbourg, France
e-mail: peter.dunne@applied-magnetism.com

© Springer Nature Switzerland AG 2021 1593


J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_35
1594 P. Dunne

is covered, with an overview of new approaches to wastewater treatment and


biological matter isolation using magnetic nanoparticles.

Introduction

The ability of magnets and magnetic fields to exert force and induce motion is
probably their most striking property. The forces may be exerted on currents of
moving electric charges or on preexisting or induced magnetic moments. Much of
this handbook is concerned with electronic currents in solid conductors. Here the
focus is on currents associated with chemical ions or free radicals. These are often
in solution, which leads us into the areas of electrochemistry and hydrodynamics.
In the early nineteenth century, both types of currents were being explored,
and the contributions of Michael Faraday were preeminent. He visualized the
magnetic field by lines of force and demonstrated the principle of the electric motor
with a current-carrying conductor. He also pioneered electrochemistry and in 1832
launched the field of magnetohydrodynamics (MHD) when he tried to measure the
potential difference across the Thames induced by its motion through the Earth’s
magnetic field. The attempt failed because his equipment was not sensitive enough;
the river was not sufficiently conductive, and the river bed short-circuited the
signal [1].
Our other topic, magnetic separation, has a longer history, dating from the first
patent in 1792, by William Fullerton, for the separation of iron minerals by means of
a magnet [2]. Between then and the start of the twentieth century, there was a gradual
flowering of the technique, with more than 300 patents awarded in the United States
by the end of the nineteenth century [2], for a huge variety of different types of
industrial magnetic separators. We first recall, how magnetic fields are produced
and how materials, especially liquids, respond to them.

Magnetic Fields

Magnetic fields have two sources; the first is the motion of electric charges, as
described by Ampère’s law
∇ × B = μj (1)

where μ is material permeability and j is the current density, which leads to the
generation of a magnetic field B. The second source of magnetic fields is the
intrinsic magnetism of elementary particles, especially the electron in condensed
matter. The magnetic moment of an electron is

ms = −gS μB S/h̄ (2)

where gS is the spin g-factor ≈2, μB is the Bohr magneton, and S is the spin angular
momentum of the electron. For an atom, the spin angular momentum S is the sum of
33 Magnetochemistry and Magnetic Separation 1595

all individual spin momenta; likewise the atomic orbital angular momentum L is the
sum of individual orbital momenta, and the total angular momentum is J = L + S.
Then the magnitude of the dipole moment of a free atom in space is

matom = gJ μB J (J + 1) (3)

where gJ is the Landè g-factor

J (J + 1) − L(L + 1) + S(S + 1)
gJ ≈ 1 + (4)
2J (J + 1)

and for a spin-only moment, the magnetic moment becomes



mspin-only = gS μB S(S + 1) (5)

The contribution of the spin and orbital angular momenta of the atom combine to
give the magnetization M, which is measured in some direction (conventionally the
z-direction). M is then defined as

N
M= mz (6)
V

where N is the number of atomic magnetic moments m of the sample and V is the
sample volume. Finally in the Sommerfeld convention, a third field, the auxiliary
field H , is related to the other two by

B = μ0 (H + M) (7)

where μ0 is the permeability of free space and M is the volume magnetization. In the
absence of an external magnetic medium, the H -field lines are identical to B-field
lines outside a magnet (the stray field) and only differ by a factor μ0 , (B = μ0 H );
but inside they point in a different direction. B is solenoidal (∇ · B = 0), meaning
that its field lines always form closed loops, or extend to infinity, whereas H is
conservative and irrotational (∇ × H = 0), provided no electric currents j are
present. Only in this case can H be represented by sources and sinks similar to the
electric field E (Fig. 1).

Magnetic Susceptibility

In the magnetic systems of interest here, the magnetization M is usually pro-


portional to the local value of H . If a magnetic field H of a certain intensity
is applied to a substance, the intensity of the external magnetic field within that
substance may be higher or lower than in the surrounding space. In the first case,
the substance is diamagnetic, and in the second, it is paramagnetic. In ferromag-
1596 P. Dunne

Fig. 1 H , M, and B fields for a uniformly magnetized material in the absence of an external
magnetic field. (Adapted from [3])

nets and ferrimagnets, the intensity of the field may be increased a millionfold.
Paramagnetic materials commonly contain transition metal elements or lanthanides.
Diamagnetism, however, is a universal property of matter. All condensed matter has
an underlying diamagnetism which may be masked if paramagnetism is also present.
For most substances, except ferromagnets, their initial response (i.e., magnetization
M) is proportional to the externally applied magnetic field H

χ = M/H (8)

where χ is the dimensionless magnetic susceptibility. A diamagnetic material, with


χ < 1, is weakly repulsed by a magnetic field gradient. It can be thought of as an
example of Lenz’s law, where currents of electrons around their respective nuclei,
induce a moment to oppose the external magnetic field. Diamagnetic susceptibilities
are very small, χ = −9 · 10−6 for water, and even for the most diamagnetic
materials, bismuth and pyrolytic graphite, susceptibilities are only −1.66 · 10−4 and
−4·10−4 , respectively [3]. The exceptions are superconductors. When the Meissner
effect is complete for a type I superconductor, or a type II below Hc1 , the material
behaves as a perfect diamagnet as it has susceptibility χ = −1 [3]. In contrast,
paramagnetic materials are weakly attracted by external magnetic field gradients
and have a positive magnetic susceptibility (χ > 1). When the susceptibility is
small, the demagnetizing field can be neglected. Upon removal of the external
field, paramagnets lose their magnetization due to thermal motion randomizing
the magnetic moments. For localized atomic moments and small external magnetic
fields, the temperature dependence of magnetic susceptibility follows the Curie law

C
χ= (9)
T
33 Magnetochemistry and Magnetic Separation 1597

Table 1 Molar susceptibility and susceptibility of 1 molar solutions of selected paramagnetic 3d


ions in water. (Adapted from [12])

χm × 10−9 (m3 mol−1 ) χsoln × 10−6 (1 mol)


S exp exp exp
3dn Ion S peff peff
1 Ti3+ ,V4+ 1/2 1.73 1.7 15.2 6.2
2 Ti2+ ,V3+ 1 2.83 2.8 41.3 32.3
3 V2+ ,Cr3+ 3/2 3.87 3.8 76.1 67.1
4 Cr2+ ,Mn3+ 2 4.9 4.9 126.6 126.6
5 Mn2+ ,Fe3+ 5/2 5.92 5.9 183.5 174.5
6 Fe2+ ,Co3+ 2 4.9 5.4 153.7 144.7
7 Co2+ ,Ni3+ 3/2 3.87 4.8 121.5 112.4
8 Ni2+ 1 2.83 3.2 54.0 44.9
9 Cu2+ 1/2 1.73 1.9 19.0 10.0

 
where C = μ0 μ2B ng 2 J (J + 1) /3kB is the Curie constant, n is the number
density of magnetic moments, g is the Landé g-factor, μB is the Bohr magneton,
and kB is Boltzmann’s constant. Two frequently used quantities are the mass
susceptibility χmass = χ /ρ with units m3 kg−1 where ρ is density; and the molar
susceptibility χmol = MW χmass = MW χ /ρ with units m3 mol−1 where MW is
molecular density. Rewriting the molar Curie law susceptibility for non-interacting
spin-only moments, it becomes [3] (see Table 1)

μ0 NA g 2 μ2B S(S + 1)
χm = (10)
3kB T

where NA is Avogadro’s number, g is the g-factor (=2 for a spin only moment),
S is the ion spin angular momentum, and kB is Boltzmann’s constant. This can be
simplified to

1.571 · 10−6 peff


2
χm = (11)
T

S = g S(S + 1) is the effective spin Bohr magneton number. Table 1
where peff
shows the calculated and measured magnetic moments for 3d spin-only moments.
S is the good quantum number for 3d ions in compounds, predicting values close
to measured values. For this to occur, the orbital angular momentum L must be
quenched. This situation arises in condensed matter where the ligands generate
a non-spherically symmetric electric field that stabilizes orbitals with no angular
momentum. The time-averaged orbital angular momentum is zero. Conversely for
4f ions, the
√ total angular momentum J is the correct quantum number to use, and
peff = gJ J (J + 1). The spin-orbit coupling in 4f ions is much larger than in
3d ions (four to ten times as large), and their ligand fields are three to ten times
smaller [4], so quenching is not as effective in 4f ions (Table 2).
1598 P. Dunne

Table 2 Molar susceptibility and susceptibility of one molar solution of selected paramagnetic
4f ions in water. (Adapted from [12])
exp exp exp
4fn ion S L J gJ p0 peff peff χm × χsoln
10−9 ( m3 mol−1 ) 10−6 (1 mol)
1 Ce3+ 1/2 3 5/2 6/7 2.14 2.54 2.5 32.9 23.9
2 Pr3+ 1 5 4 4/5 3.2 3.58 3.5 64.6 55.5
3 Nd3+ 3/2 6 9/2 8/11 3.27 3.52 3.4 60.9 51.9
4 Pm3+ 2 6 4 3/5 2.4 2.68 - - -
5 Sm3+ 5/2 5 5/2 2/7 0.71 0.85 1.7 15.2 6.2
6 Eu3+ 3 3 0 0 0 0 3.4 60.9 51.9
7 Gd3+ 7/2 0 7/2 2 7 7.94 8.9 417.6 408.5
8 Tb3+ 3 3 6 3/2 9 9.72 9.8 506.3 497.3
9 Dy3+ 5/2 5 15/2 4/3 10 10.65 10.6 592.3 583.3
10 Ho3+ 2 6 8 5/4 10 10.61 10.4 570.2 561.1
11 Er3+ 3/2 6 15/2 6/5 9 9.58 9.5 475.8 466.7
12 Tm3+ 1 5 6 7/6 7 7.56 7.6 304.5 295.4
13 Yb3+ 1/2 3 7/2 8/7 4 4.53 4.5 106.8 97.7

Magnetic Susceptibility of Liquids

The magnitude of the dimensionless susceptibility, |χ |, is typically small in


magnetochemical or electrochemical systems (10−6 − 10−1 ) and χ is written as
the sum of the susceptibility of the solute and ions in solution. For water, this is:

χ = χ water + χm c (12)

where χ water = −9 · 10−6 , χm is the molar susceptibility, and c is the molar


concentration of dissolved ions.
Measured ionic moments have been shown to agree with the spin-only moments
of the ions rather than the free ion moments. They are independent of concentration
[5, 6, 7, 8, 9, 10]. Initially this may seem strange, since the ions are fully dissolved
and free to move in solution. However, the ions are solvated by water molecules.
Due to the electronegativity of the oxygen in water, a solvation shell forms around
each ion, and for 3d ions, hexa-aquo complexes are formed [11], meaning the ion
sits at an octahedral site, where the ligand field quenches L. The susceptibility of
one molar solutions of 3d and 4f ions are included in Tables 1 and 2.

Pascal’s Constants

As previously mentioned, diamagnetism is a universal property of matter. Often to


accurately determine the magnetic susceptibility of a paramagnetic compound, a
diamagnetic correction must be applied to any measured data (χP = χmeas + χD ).
33 Magnetochemistry and Magnetic Separation 1599

Pascal was able to conclude from a vast number of measurements that the
diamagnetic susceptibility of a molecule, χD , could be determined by summing the
susceptibility of every constituent atom, χDi , and bond, λi [13, 14]:
 
χD = ni χDi + λi (13)
i i

This is only an empirical law, yet its success at estimating χD remains unchallenged,
allowing the determination of susceptibilities where no other method is available,
such as in triarylmethyl radicals which have significant diamagnetic susceptibilities,
but without a direct method to determine them [13]. These constants are tabulated
in [14].

Magnetic Field Effects in Chemistry

Since the 1970s, weak magnetic fields have been known to strongly influence certain
chemical reactions [15], which was surprising because the energies involved are
<0.1 meV. The reactions in question involve the simultaneous production of two
intermediate, short-lived radicals. Each radical has at least one unpaired spin, but
because they are produced simultaneously, spin multiplicity is conserved, and their
spin states are correlated in an antiparallel S = 0 singlet state ‘S’, or parallel S = 1
triplet state ‘T’. There are three ways to arrange the triplet state

1
|T0  = | √ (↑ + ↓)
2
|T+1  = |↑↑
|T−1  = |↓↓ (14)

but only one way to arrange the singlet state.

1
|S = | √ (↑ − ↓) (15)
2

The unpaired electrons tend to be delocalized [16, 17] and interact with many
nuclei in the radicals. Nuclei with odd numbers of protons, neutrons, or both are
weakly paramagnetic, having magnetic moments which generate local hyperfine
magnetic fields. Most isotopes with paramagnetic nuclei are rare, but two, 1H and
14N, have almost 100% natural abundance and are common in organic radicals [16].

For radical pairs which have these hyperfine fields, their spin states are not static,
rather a fraction of radicals oscillate between singlet and triplet states at frequencies
dependent on the hyperfine field strengths found in the radical and can be of order
1 mT for 14N containing radicals [16]. This singlet-triplet (S-T) interconversion
can be driven not only by these nuclear spin states but additionally by an external
1600 P. Dunne

magnetic field. Due to the Zeeman interaction, the electron spin precesses at the
Larmor frequency around the external field, which is about 1.4 MHz in the Earth’s
magnetic field (50 μT) [16].
Figure 2b is a schematic of the simplest radical pair scheme. Radicals are
commonly formed by photochemical excitation, where the absorption of a photon
results in the transfer of an electron from a molecule A to molecule B in close
proximity, or to different parts of a single molecule. After this, an applied magnetic
field μ0 H can regulate the interconversion between S and T states, which then
decay to their singlet or triplet products. The simplest examples being a return to
the neutral AB reactants for the singlet state and different product, C, for the triplet
state. If the external magnetic field favors interconversion of S to T, then more C will
be produced, while if the field suppresses interconversion, less C will be produced,
and more radical pairs can recombine to AB.

a)
Electron
spins European Robin

Retina

B
Nuclear
spin

b)

– Singlet
N products
hO S ●+ ●–
[AB] [A B ]
Singlet AB
+ μ 0H

T
[A●+ B●–]
Triplets
T+ T–
Electron spin Triplet
N products

Nuclear spin T0 + C

Fig. 2 Schematic of radical pair production and how it is used to explain avian magnetoreception.
(Adapted from [18])
33 Magnetochemistry and Magnetic Separation 1601

Without the hyperfine fields, there would be no S-T interconversion in such a


weak magnetic field, so the question arises, how is it possible for terrestrial fields to
influence S-T interconversion? The energy of an unpaired spin in 50 μT is orders
of magnitude smaller than thermal energy, i.e., kB T, so the spin states cannot be
in equilibrium, yet there is unequivocal evidence for their effect on the course of
reactions [15,19,18,20]. Indeed, the magnetic field effect on radical pair reactions is
observable because their chemistry is largely determined by their spin states, which
are non-equilibrium states thermally isolated from their environment. It can take as
much as 1 μs for their spin states to reach equilibrium, allowing weak interactions
like terrestrial magnetic fields, to influence chemical pathways [16]. In addition,
hyperfine fields are highly anisotropic, so the relative yield reaction product is also
dependent on their angle with respect to the direction of the external field, a key
requirement for a magnetic compass.
Radical pairs and dynamic spin-triplet conversions are readily observed in
nuclear magnetic resonance (NMR) spectroscopy and electron paramagnetic res-
onance (EPR) spectroscopy [16], so it was widely expected that spin-triplet
conversion effects might be detected in biological systems. Radicals are widespread
in biology, playing a role as transient reaction intermediates in various enzyme
reaction cycles [21], and having a key function in activating signaling pathways
that affect gene expressions and initiate cell death [22, 23]. However, many of
the reported magnetic effects on a variety of enzyme systems have not been
independently reproduced [24, 25, 26, 27, 28, 29, 20]. One area where there has
been some progress is with the photoactive protein, cryptochrome, which forms
radical pairs in vitro after excitation by blue light [15, 30, 31, 32]. This is
important in the understanding of avian navigation, and after nearly 40 years of
active research, cryptochrome magnetoreceptors [33, 19, 18] in the retina of birds
(Fig. 2) are still the most likely candidate to explain the phenomenon of avian
magnetoreception [19, 32, 17, 16].

Magnetoelectrochemistry

A magnetic field was shown to influence the corrosion of an iron electrode in


1885 [34], which the effect was only explained relatively recently in terms of MHD
effects [35, 36, 37]. It was not until the 1930s that MHD-like effects were studied
by Hartmann and Lazarus in liquid metals [38]. Then in 1947, Alfvèn coined the
term magnetohydrodynamics and a new field was born [1]. Progress in magneto-
electrochemistry was sporadic before the 1970s. The earliest example is a study
carried out by Yang in 1954 [39], who observed an increased roughness under
magnetic fields of iron, nickel, and cobalt deposits along ridges on the substrate
formed by polishing. He attributed the effect to concentration of magnetic field
lines close to protruding ridges, creating a magnetic field gradient, which then
acted to attract the paramagnetic ions. It was not until the 1970s that Aogaki
et al. analytically derived a set of equations relating the limiting current in an
electrochemical cell to the applied magnetic field for conduit geometries such as
1602 P. Dunne

shown in Fig. 5 [40, 41]. These are the classical papers for the often surprising
science of magnetoelectrochemistry.

Electrochemistry

We first summarize some basic ideas in the absence of a magnetic field. In the
space of a few microns, electron kinetics, diffusion, and hydrodynamics combine
at the electrode-solution interface to determine much of what is important in
electrochemistry. The are two distinct interfacial regions, the diffusion layer and
the double layer [42]. The former is a zone 1–100 μm thick, where ion diffusion is
driven by a concentration gradient ∇c that is influenced by convection. The latter is
a layer a thousand times thinner where huge electrostatic fields control the electrode
kinetics in a region that is sensitive to surface tension and the structure of the local
charge distribution. The double layer is further divided into a compact charged layer
usually 0.5–10 nm thick coupled to a broader diffuse layer.
Even for simple one-electron transfer reactions, a number of steps are required
to oxidize or reduce an ion from solution at the electrode, requiring classical or
quantum mechanical descriptions [43,44,45,46] (see [47] for a rigorous description
of electron transfer including second-order effects). First, the ion must be driven
to the electrode surface by mass transport, which must be at a thermodynamically
favorable potential for the reaction to take place and have reasonably fast kinetics to
sustain it (Fig. 3). The charge transfer current is strongly dependent on the electrode
potential and the concentration of relevant species at the electrode surface, being
described by the Butler-Volmer equation when under kinetic control [42]
    
αc zFη αa nFη
j = j0 exp − − exp (16)
RT RT

Electrode surface region Bulk solution


Electrode Chemical
reactions Mass transfer
Oxsurf Oxbulk
Ox´
Desorption
Adsosption
Ox´ads

ne

Red´ads
Desorption
Chemical
Adsosption reactions
Red´ Redsurf Redbulk

Fig. 3 Simple schematic of electron transfer at an electrode. (Adapted from [48])


33 Magnetochemistry and Magnetic Separation 1603

where αc and αc are the cathodic and anodic charge transfer coefficients, z is the
number of electrons involved in the reaction, F is Faraday’s constant, R is the gas
constant, η is the overpotential, and j0 is the exchange current density. In more
detail, η = E − Eq is the potential difference between the measured electrode
potential E and the thermodynamically determined potential of the reaction Eq ;
and the exchange current density j0 is the current when there is no net electrolysis
and the electrode is at equilibrium, i.e., ja + jc = 0.
Three processes govern the mass transport of ions, diffusion – movement of
a species due to a concentration gradient; migration, the movement of a charged
species in an electric field; and advection, stirring or other bulk hydrodynamic
transport of the medium, through density gradients or external forces. Confusingly,
convection is sometimes used as a synonym for advection, but convection is actually
defined as the collective motion of all particles in a fluid.
Mass transport can be described by the Nernst-Planck equation [42]

zi F
J i = −Di ∇ci − Di ci ∇φ + ci v (17)
RT

where J i is the flux of species i ( mol s−1 m−2 ) at a given position from the
surface, Di is the diffusion coefficient m2 s−1 , ∇ci is the concentration gradient,
zi and ci are the charge (dimensionless) and concentration (mol m−3 ) of species
i, ∇φ is the potential gradient, and v is the fluid velocity of a volume element in
solution. Migration is typically insignificant in electrode currents so long as an inert
supporting electrolyte is introduced to the solution, which can carry the migration
part of the current, without making any contribution to the reaction current, while
convection is the motion of a fluid under the action of a force. There are two
kinds of convection: the first is natural convection which arises from thermal and
density gradients and is common to all solutions. The second is forced convection
which is due to an external force, such as provided by mechanical stirring or
magnetohydrodynamic stirring. The velocity field in a fluid can be described by
the Navier-Stokes equation, where ρ is the fluid density

Inertia (per volume)



Divergence of stress

 ∂v 
ρ + v ·
∇v = −∇p + μ∇ 2 v + f (18)
∂t




Convective Pressure Viscosity Other
Unsteady acceleration gradient body
acceleration forces

Two kinds of fluid flows most relevant to electrochemistry are laminar flow and
turbulent flow. In a laminar flow, the fluid velocity is smooth and steady and occurs
in separate layers or laminae. A flow is turbulent when the fluid motion is chaotic
and unsteady, with only an average net velocity in one direction. The most important
dimensionless numbers used in fluid dynamics are listed in Table 3 (Fig. 4).
1604 P. Dunne

Table 3 Characteristic dimensionless numbers used in fluid dynamics, where v is the fluid
velocity (m s−1 ), l is the characteristic length scale (m), e.g., pipe length, ν is the kinematic
viscosity (m2 s−1 ), D is the diffusion coefficient (m2 s−1 ), kd the convective mass transfer
coefficient (m2 s−1 ), and δ is the diffusion layer thickness (m). Note that convection is the total
collective motion of a fluid advection is fluid transport by bulk motion only
Number Formula Description
Reynolds number Re = vl/ν Measure of the ratio of inertial to
viscous forces. For small Re, flows
are laminar, and for large Re, flows
become turbulent
Schmidt number Sc = ν/D Ratio of momentum diffusivity (vis-
cosity) to mass diffusivity. Relates
δH and mass boundary layer
Peclet number Pe = lν/D(= Re.Sc) Ratio of advection of a physical
quantity of a fluid, e.g., mass or heat
transfer by flow to its rate of diffusion
Sherwood number Sh = kd l/D = 1/δ Ratio of convective mass transport to
diffusive mass transport
Hydrodynamic layer thickness δH = (ν/D)1/3 δ A characteristic length which con-
tains most of the fluid velocity gra-
dient between a bulk flow and fluid-
solid interface

Fig. 4 A three-electrode
electrochemical cell, I
- +
consisting of a working
electrode (WE), reference
electrode (RE), and counter
electrode (CE) Vcell

WE
RE CE

A standard electrochemical cell consists of a working electrode, upon which the


reactions of interest are studied, a counter electrode to supply the current, and a sta-
ble reference electrode that is used to measure the potential at the working electrode.
Two common conflicting conventions are used in electrochemical literature. In the
older American convention, anodic (positive) potentials are plotted to the left of the
origin, and cathodic currents are defined as positive. In the European convention,
adopted by IUPAC, anodic (positive) potentials are plotted to the right of the origin,
and anodic currents are positive. Only the IUPAC convention will be used here.
33 Magnetochemistry and Magnetic Separation 1605

Magnetic Forces in a Uniform Magnetic Field

In the Absence of Currents


In a uniform magnetic field, H , the force density acting on a non-uniformly
magnetized material in the absence of currents is

F = μ0 [∇ · (H + Hd )]H (19)

where H d is the demagnetizing field. However, the applied field H is uniform, and
as long as the demagnetizing field is negligible, there is no net force, which is true
for aqueous electrochemical systems, but not in ferrofluids (see section “Magnetic
Forces in a Non-uniform Magnetic Field”). There is also a weak stress in elec-
trolytes, known as the Maxwell stress, which in tensor form for an applied magnetic
field B is [1]

1 1 2
σij = Bi Bj − B δij (20)
μ0 2μ0

where 0 is the permittivity of free space and δij is Kronecker’s delta, or simply σ =
MB where the units are N m−2 [49]. This is known to deform volumes containing
paramagnetic ions, such as paramagnetic tubes [50].

Lorentz Force
In aqueous solutions, moving charges interact with a magnetic field through the
Lorentz force

FL = j × B (21)

Acting as a driving force, the Lorentz force induces convection in a bulk solution
[40, 41]. An example is the Aogaki cell (Fig. 5) which consists of two parallel plate
electrodes embedded in the wall of a tube, with a perpendicularly applied magnetic
field which induces a flow in the tube. The limiting current densities are described
by

γ
jl = j0 + αB β c0 (22)

where j0 is the zero field current density, c0 is the ion concentration, α is a


constant that depends on the cell configuration, and β and γ are indices dependent
on the flow profile of the cell. For a narrow channel in a fully developed flow
β = 1/2 γ = 3/2, whereas β = 1/3 γ = 4/3 for a wide channel. This power
law dependence is a universal behavior with many other redox systems such as
[Fe(CN)6 ]−3/−4 exhibiting the same relationship [51] and has also been validated
by FEM simulations [52].
1606 P. Dunne

Fig. 5 The Aogaki cell [53], to potentiostat


(a) Schematic configuration;
a: working electrode, b: f
counter electrode, c: Luggin
capillary, d: solution flow, e:
magnetic field, and f: vessel e b
filled with solution. (b) MHD d
main flow and boundary
layers; a: Luggin capillary, b:
working electrode, c: counter
c
electrode, d: diffusion layer,
e: hydrodynamic boundary a
layer, and f: streamlines

Z X a
Y b
f d
e
B
f
i
f e
d
c

x
0 x1 x2 l

Characteristic MHD Numbers


Similar to section “Electrochemistry”, where we introduced convection and the
dimensionless numbers used in fluid dynamics, there are dimensionless numbers
specific to magnetohydrodynamics that are used as indicators of the relative
interaction strengths. The first of these is the Magnetic Reynolds number

Rm = ul/λ (23)

where u is characteristic velocity, l is the characteristic length scale, λ is the


magnetic diffusivity λ = 1/μ0 σ , and σ is the conductivity of the medium. In
contrast to the other dimensionless numbers, the magnetic Reynolds number has
nothing to do with forces, rather it gives an estimate of the advection of a magnetic
field due to the motion of a conductive medium versus magnetic diffusion. It can
also be viewed as a measure of the dimensionless conductivity. When Rm is large,
diffusion is weak, and the magnetic field lines act like elastic bands frozen into the
conducting medium due to Maxwell stresses. This is the case, for example, in the
huge length scales of interstellar media such as astrophysical plasma (Rm > 106 ),
33 Magnetochemistry and Magnetic Separation 1607

and in Earth’s liquid outer core (Rm ≈ 103 ), and whose motion is sufficient to
sustain a magnetic field against dissipation [1]. In aqueous solutions, which are poor
conductors, Rm is small (≤1) and the magnetic field induced by motion of the liquid
is negligible compared to the imposed field. The generation of a magnetic field due
to the motion of a conductive liquid has only been achieved in large experiments
involving liquid sodium or mercury [54]. The second commonly used dimensionless
number is called the interaction parameter

Nm = σ B 2 l/ρu = l/uτ (24)

where τ = (σ B 2 /ρ)−1 is the magnetic damping time. The interaction parameter is


the ratio of Lorentz forces to inertia. Finally, there is the Hartmann number
 
Ha = Nm Re = Bl σ/ρν (25)

which squared (Ha2 ) represents the ratio of Lorentz forces to shear forces. Previ-
ously MHD effects in solution were observed to have a dominant role and were
the focus of a number of reviews, which mostly concentrate on either deposition or
dissolution of metals [55, 56, 57, 58]. When the magnetic field is parallel to the
electrode surface, i.e., nominally perpendicular to the current, the Lorentz force
acts parallel to the electrode surface, generating a fluid flow orthogonal to both the
current and magnetic field (Fig. 6a). If the magnetic field is perpendicular to the
surface (Fig. 6b), and parallel to the current, there can still be a Lorentz force due
to non-collinear current lines at the edge. This can generate vortices at an electrode
unless care is taken to minimize edge effects. A further development to this case is
when secondary micro-MHD vortices appear at an electrode (Fig. 6c). These arise
from protuberances on the electrode surface, causing symmetric and asymmetric
concentration fluctuations close to the electrode surface [53]. Asymmetric fluctua-
tions are triggered in the double layer, and determine the reaction rate, yielding 2D
nuclei, and symmetric fluctuations are triggered in the diffusion layer, determining
the reaction rate of each crystal/grain and forming 3D nuclei [59].

Deposition
During metal electrodeposition, applied magnetic fields have been shown to influ-
ence the surface morphology [39, 55, 56, 61, 62, 63, 57, 64, 65, 66, 67, 68, 69, 70,
71, 72, 73, 74, 75, 76, 77, 78], crystal structure and orientation [79, 64, 80, 81, 82, 83],
deposition rate [84,85,86,87,88,89,90,65,91,92,66,67,69,73,93], alloy composition
(NiCu [72], CoNi [94, 95], NiRe [96]), magnetic properties [97, 64, 72], surface
adhesion [98], mechanical hardness [99], deposit crack density [100], and improved
corrosion resistance [101]. The surface morphology and crystal orientation depend
not only on the magnitude of the magnetic field but also its orientation, as set
out in Fig. 6. Due to the different local MHD flows, the surface of metal deposits
such as CoFe can be smoother with B parallel to the electrode surface, or rougher
for B perpendicular to the electrode [102, 78], or also with compounds like
1608 P. Dunne

Fig. 6 Hydrodynamic flow at an electrode in a uniform magnetic field. (a) Field applied parallel
to the electrode surface. The primary MHD flow is parallel to the surface. (b, c) Field applied
perpendicular to the electrode surface. In (b) the primary MHD flow is a vortex around the rim but
in (c) secondary micro-MHD vortices arise around protuberances on the surface. (From [60])

hydroxyapatite, with plate-like crystallites without a magnetic field, which become


needlelike, or spherical for B perpendicular or parallel to the electrode [98]. Lorentz
force stirring can also be used to improve deposit homogeneity onto complex
geometries such as cylinders [103] because they are complex flows that are three
dimensional in nature [104, 105, 106, 107].
Optical velocity profile measurements during electrodeposition have shown that
the primary MHD flow only acts during the initial stages of electrodeposition [108,
109]. At long times, buoyancy forces are greater, or equal, to the Lorentz force,
and a stable diffusion layer is formed. This fits well with the observation that the
early stages of Cu [53], Cu2 O [83, 110, 111], CoFe [112], Fe [113, 112], and
hydroxyapatite [98] nucleation are sensitive to the hydrodynamic stirring of an
external magnetic field. However, with careful design of a non-uniform magnetic
field, the stability of the MHD convection can be enhanced and has been observed
in both experiments and simulations [114]. In contrast, the secondary micro-
MHD flows inside the diffusion layer continue to act throughout the deposition
process [53]. Other experimental evidence for the micro-MHD effect include the
modification of fractal geometry of Zn and Cu deposits [61, 87, 68], complex Cu2 O
cube morphologies (Fig. 7), and in the morphology of Cu, Co, CoNiP, Ag, and Zn
thin films [73, 80, 115, 76, 116].
Furthermore, the micro-MHD effect has led to the study of vortex chirality during
copper deposition [117, 118], and there is now evidence that the imposed chirality
on a metal surface during deposition in a magnetic field can lead to it being sensitive
subsequently to the chirality of organic molecules in the absence of a magnetic field
[119,120]. The effect of magnetic field orientation on the chirality of electrodeposits
can be clearly seen in Fig. 8, for the change in Zn dendrite chirality for a magnetic
field out of or into the plane of the page.
Metal deposition is not the only reduction process to be probed with magnetic
fields; both gas evolution and organic compounds have been used to further the
understanding of MHD effects. It has become quite clear that gas evolution due
to the production of hydrogen, oxygen, or carbon dioxide is indeed sensitive to
33 Magnetochemistry and Magnetic Separation 1609

Fig. 7 SEM images of Cu2 O crystals grown under magnetic fields A1–A3: 0 T; B1–B3: 6 T;
C1–C3: 12 T. (From [111])

Fig. 8 Zinc electrodeposits grown from 0.1 M ZnSO4 at 10 V in a flat horizontal cell in the
presence of a 0.35 T magnetic field applied (a) vertically upward and (b) vertically downward.
(From [68])

magnetic fields. When the Lorentz and buoyancy forces sum, the mechanical effect
of the Lorentz force, whether it is to sweep the surface or apply torque to the bubbles,
will tend to dislodge the bubbles sooner than would otherwise be the case [121].
Whereas the magnetic field can have a stabilizing effect on the bubbles when the
Lorentz force opposes buoyancy. The average bubble size should be reduced by
the field in the former case and increased in the latter. This is indeed observed
during the deposition of CoFe or Cu [122,123,124,125] and has also been observed
during the direct electrolysis of water [126,127,128,129,130]. This effect is already
1610 P. Dunne

found to be efficacious to direct methanol fuel cells, with a 12.5% improvement in


maximum power density, due to the Lorentz force acting on CO2 bubbles [131].
Systematic studies of MHD effects on organic compounds have also been
carried out showing a very clear MHD effect using electromagnets and permanent
magnets [132, 133]. Lee et al. used neutral molecules such as nitrobenzene (NB) to
probe the effect of the Lorentz force on the products and showed that the limiting
current is dependent on the concentration of the reactants and products close to the
electrode. Under increasing magnetic fields, the reduced, charged, and paramagnetic
product NB•− is swept away from the electrode, being replaced by NB, thereby
increasing the limiting current [134]. Going beyond MHD effects, which act on
the outer diffusion layer, at low to intermediate uniform magnetic fields, can exert
a large and quite unexpected influence on the double layer – the first few nm at
the electrode-solution interface – through Maxwell stress acting on paramagnetic
species in this region [135]. A 0.5 T in-plane field during the reduction of NB
or oxidation of NB•− modifies its double layer capacitance and effective charge
transfer resistance by up to 50%, indicating a shift of the plane of closest approach
for the reduced species by up to 0.25 nm.

Corrosion
Previously mentioned was the fact that the very first magnetic field effects were
observed for the corrosion of iron. Until relatively recently it was assumed that only
ferromagnetic or paramagnetic electrodes exhibited a change in corrosion activity
under applied magnetic fields [136, 137, 138, 139, 140]. One figure of merit for cor-
rosion activity, the rest potential, is the potential difference between the immersed
surface of interest and a reference electrode. A cathodic shift of the rest potential
means an increase in corrosion, while an anodic shift means the reverse, and it is
only in the last decade that it was recognized that rest potential shift during corrosion
is due to micro-MHD effects acting on local currents [35, 36, 141, 37, 142, 143].
Interestingly, micro-MHD theory was originally developed to account for copper
corrosion patterns and behavior in uniform magnetic fields [53, 144, 145, 146].
Nevertheless, the effect can be magnified for ferromagnetic electrodes such as
Fe [147, 148, 149, 150] where the dissolved species (i.e., iron ions) remain close
to the electrode, displacing H + , thus inducing a cathodic shift of the rest potential
and a suppression of the corrosion current that can be enhanced even further by
magnetic field gradients (see the section below). Typical rest potential shifts are on
the order of −100 mV/T [37].

Magnetic Forces in a Non-uniform Magnetic Field

If a non-uniform magnetic field is applied to an electrochemical cell, another


magnetic force can act on the fluid, namely, the Kelvin force

F = μ0 (M · ∇)H (26)
33 Magnetochemistry and Magnetic Separation 1611

which can also be written as the magnetic field gradient force

F ∇B = (1/2μ0 )χm c∇B 2 (27)

It must be noted that the Kelvin force is only valid for linear media where the
susceptibility is linear in density, χ = αρ, where α is a proportionality constant
[151], and where we have assumed ∇ × M = 0. This is not true if we cannot
neglect the demagnetizing field H d = Nχsoln H , which is the case in ferrofluids of
order 10−1 times the applied field, in contrast to solutions used in electrochemistry
with much lower susceptibilities and demagnetizing fields of order 10−4 to 10−5 .
The correct description of the force in the absence of currents is then the Helmholtz
force rather than the Kelvin force [152, p. 127]

   
μ0 ∂χ μ0 2
F = ∇ H ρ
2
− H ∇χ (28)
2 ∂ρ T ,P 2

In this formulation, there is also a concentration gradient force, but it is only a


higher-order correction term to the force density.
In linear isotropic media, the susceptibility is proportional to the density,
ρ(∂χ /∂ρ) = χ , and the Helmholtz force reduces to the Kelvin force F =
(μ0 /2)χ ∇H 2 = μ0 M∇H  , meaning there is no concentration gradient force in
dilute electrochemical systems. However, using an incorrect formulation of the
magnetostatic energy for an electrolyte, it is possible to arrive at an erroneous
first-order magnetic concentration gradient term [153]. It is now recognized that
many of the phenomena previously attributed to a magnetic concentration gradient
force [154, 155, 91, 156, 157] are in fact due to edge effects in three dimensional
Lorentz force convection [158, 159, 106, 160, 114, 161, 162].
The field gradient force is a well-known body force, whose most dramatic
manifestation is in magneto-Archimedes levitation. The difference in susceptibility
of an object and its environment plus the buoyancy differences can lead to levitation
in ferrofluids [163], paramagnetic liquids [164, 165], liquid oxygen [166], or
directly in air [167, 168, 169, 170, 171, 172, 173]. An example of levitation is shown
in Fig. 9. The field gradient force, like the Lorentz force, is a body force acting
across a volume of a solution, whereas diffusion normally acts on individual ions
or molecules in solution. Therefore one would expect the field gradient force to
influence convection and not diffusion. Taking typical values in an electrochemical
cell, χm = 1.9 · 10−8 m3 kg−1 , c = 1 mol l−1 , B = 1 T, and ∇B = 10 T m−1 , and
F∇B = 152 N m−3 (see Table 4), which is a moderate force density, but the energy
of a Co2+ ion (S = 1/2) in a 1 T field is

μ0 g 2 μB S(S + 1)B 2
E= = 7 · 10−7 eV
2kB T
1612 P. Dunne

Fig. 9 From top to bottom,


C, Si, and Ti levitating in a
2.67 M Dy(NO3 )3 solution in
the fringing field of a 1.5 T
electromagnet [165]

which is considerably less than the thermal energy driving diffusion (kB T ∼
1/40 eV). Unlike the Lorentz force, the field gradient force is normally conserved,
meaning it cannot generate convection. However, it is not conserved if (i) there is a
non-uniform distribution of paramagnetic species and (ii) the concentration gradient
and field gradient are non-collinear allowing convection to develop [174,175]. These
two criteria can be summed up by taking the curl of the magnetic field gradient force,
which must be nonzero if it is to induce convection:
 
∇ × F ∇B = (1/2μ0 ) ∇c × ∇B 2 (29)

The Kelvin, or field gradient force, F ∇B , has been utilized in electrochemistry


for more than 20 years, starting with studies of the redox reaction of organic
compounds such as nitrobenzene (NB) [176, 177, 178, 179, 180], first in the fringing
field of a superconducting magnet, with the electrode orientated to minimize
Lorentz force stirring [176], and then with Fe and Pt electrodes to differentiate
the local effects due to the magnetic field gradient, and the Lorentz force. At
Fe electrodes, a focusing of the reduced species NB•− was observed, due to the
magnetic field gradient at the electrode surface [177, 178] in agreement with FEM
33 Magnetochemistry and Magnetic Separation 1613

Table 4 Typical forces, and estimated values, in a magneto-electrochemical cell


Typical valuea
Force Expression ( N m−3 )
Diffusion RT ∇c 1010
Migration zFc∇V 1010
Forced convection at a rotating disc electrode ρ(rω)2 /2δ0 105
Natural convection Δρg 102
Viscous drag ρν∇ 2 υ 102
Field gradient force μ0 χcH ∇H 102
Lorentz force J ×B 103
Amperian attraction (μ0 /4π r 2 )I 1 I 2 u2T un 10−5
Magnetic damping force σ (v × B) × B 101
for T = 298 K, c = 103 mol m−3 , δ = 10−4 m, z = 1, V = 1 V, ρ = 103 kg m−3 ,
a Calculated

d = 10−2 m,ω = 102 rad s−1 , Δρ = 10 kg m−3 , η = 10−3 N s m−2 , v = 10−1 m s−1 , B = 1 T,
χm = 10−8 m3 mol−1 , ∇B = 10 T m−1 , j = 103 A m−2 , and σ = 102 Ω −1 m−1

calculations [180]. A microfluidic system based on focusing NB•− at closely spaced


electrodes using the Lorentz force [179] was demonstrated, and since then a number
of other groups have focused on using magnetic fields and field gradients to realize
novel microfluidic devices, ranging from paramagnetic liquid tubes with unusual
near-frictionless flows generated in bulk solutions using tailored magnetic field
gradients [59, 181, 50], field gradient separation, and filtration devices [182, 183],
to more typical lab-on-a-chip devices using the MHD effect to pump, stir, and
concentrate solutions [184, 52, 104, 185, 186, 187, 188, 189, 190, 191].

Magnetic Structuring and Magnetophoresis


As well as uniform ∇B being applied across electrodes, locally varying gradient
fields have been used to selectively structure electrodeposits (Fig. 10). This includes
magnetized steel meshes [192], magnetized Fe wire arrays [193,194], and arrays of
permanent magnets [175, 195]. The magnetic field can directly impose a structure
on the deposits, which for solutions containing paramagnetic ions such as Fe 2+ ,
Ni 2+ , Co 2+ , or Cu 2+ leads to thicker and rougher deposits above regions of high
magnetic fields. This type of structuring is called direct templating. The reverse
case is also possible, inverse templating, when an electroactive but paramagnetic
species is introduced to the bath, leading to thinner deposits in regions of high
magnetic field. This has been shown to work for non-electroactive Mn 2+ introduced
to structure Bi [194], and for highly paramagnetic Dy 3+ on diamagnetic Zn 2+ ,
and also paramagnetic Cu 2+ [195]. The patterning effect does not depend on how
the non-uniform field is generated, but methods are equivalent, but the permanent
magnets have the advantage that neighboring ones can have their magnetization
anti-parallel, allowing much more intense magnetic fields and field gradients to be
produced [195]. Another approach to improve the structure quality is reverse-pulse
plating, with cycles of deposition and removal [196].
1614 P. Dunne

Fig. 10 Normal deposits obtained from a Cu 2+ bath and inverse deposits obtained from a
Zn 2+ /Dy 3+ bath, using parallel (top) or alternating (bottom) hexagonal magnet arrays, compared
with the contours of B 2 produced by the underlying arrays. (From [197])

The phenomenon can be understood in two complementary fashions, first as


a modulation of the diffusion layer thickness by a quasi-equilibrium between
magnetic pressure

Pm = (1/2μ0 )χmol c0 B 2 (30)

and buoyancy forces [195]. The second interpretation is that of a fluid flow driven
by the vorticity of the Kelvin force acting on a paramagnetic species in regions of
high magnetic fields. Both models capture the essence of the phenomena and are
equivalent interpretations [198, 197].
Even in the absence of a current, paramagnetic ions such as Cr 3+ , Mn 2+ ,
Co 2+ , Ni 2+ , and Cu 2+ can move along magnetic field gradient contours in silica
gels [199]. Using Mach-Zehnder interferometry, this concentration change can also
be measured directly for Mn 2+ , Dy 3+ , Gd 3+ , or Er 3+ in solutions [200, 201, 202].
Placing a magnet close to a solution containing Mn 2+ (Fig. 11), there is an
enrichment of Mn close to the magnet, which then decays after the removal of the
magnet, clearly showing the existence of a magnetophoresis effect on the solvated
ions. This implies that the ions are not acting independently but respond coherently
to the magnetic field.

Magnetic Water Treatment

Here we focus on the effects of magnetic fields on water itself. Since the 1950s, there
have been a profusion of permanent magnet devices on the market claiming to have
an effect on the formation of limescale from hard of water [203]. Claimed benefits
include being able to remove old scale and prevent new scale from forming in water
pipes or water boilers if the water passed through a magnetic field created by the
permanent magnet device. Later reports, such as by Kronenberg [204], reported
a change in the morphology of crystals grown from the mineral content in water
33 Magnetochemistry and Magnetic Separation 1615

a 1200 s b 1206 s
0 0 15

10 10

5 5
z (mm)

z (mm)
5 0 5 0

−5 −5

−10 −10

10 −15 10
0 5 10
−15

0 5 10
x (mm)
x (mm)

c 1206 s
x 10
−4

8
2
7

6
4
z (mm)

4
6
3

2
8
1

0
2 4 6 8
x (mm)

Fig. 11 (a) Enrichment of Mn 2+ in a magnetic gradient field 1200 s after the magnet was applied
on top of the cell. (b) and (c) show the concentration contour plot and the velocity vector plot
associated with the downward flow after the magnet was removed. The concentration changes in
(a) and (b) are in mM. (From [200])

which could be changed from dendrites into disk-shaped crystals with treatment of
the water, and other reported a change in the properties of calcite and aragonite in
this scale. Some of the conflicting theoretical and experimental reports include both
a decrease [205,206] and increase of surface tension [207] after magnetic treatment,
and a change in conductivity [208, 209], along with a time-dependent change of the
zeta potential and viscosity of calcium carbonate-containing water [210]. One of the
more puzzling aspects has been the report of a memory effect, with the change in
physical properties lasting for hours after the magnetic treatment [211,212,213,206,
214, 210]. The validity and reproducibility of many experimental results have been
strongly challenged [215, 216], not least because no physical mechanism could be
envisaged whereby fleeting exposure to a magnetic field gradient could influence the
nucleation of calcium carbonate from a saturated solution hours later. The situation
changed with the discovery that much of the calcium carbonate in equilibrium in
1616 P. Dunne

water was not in the form of isolated hydrated calcium cations and bicarbonate
anions but in shapeless nanometric hydrated prenucleation clusters [217, 218]. It
has been proposed that the field exposure facilitates singlet-triplet interchange of
proton dimers, which facilitates a long-term chemical modification of the clusters,
that influences subsequent carbonate nucleation [219]. The theory predicts the
design of the successful water treatment devices. Monte Carlo molecular dynamics
simulations have been contradictory in nature, showing magnetic fields to weaken
hydrogen bonds in water [220], while also suggesting the structure of water becomes
more stable with an increased number of hydrogen bonds in magnetic fields [221].
Thus, there is no clear consensus on what effects, if any, magnetic treatment of this
kind has on water.

Magnetic Separation

The final topic, magnetic separation, works on the principle of distinguishing


materials by their magnetic susceptibility, χ . As mentioned earlier, if a non-uniform
magnetic field B is imposed on a material with dimensionless susceptibility χ , the
material experiences the magnetic field gradient force density

χ
F ∇B = B∇B (31)
μ0

A classic example is that of a magnetic conveyor belt, dragging metal waste over a
series of magnets (Fig. 12), so that the trajectory of falling ferromagnetic material is
disturbed by the attractive force.
The effectiveness of magnetic separation is governed by the competition between
the magnetic field gradient and gravitational forces, particle-particle interactions,
and when in a fluid: buoyancy forces and particle-fluid interactions [223, 222, 224,
225]. Although the earliest magnetic separation applications were devoted to ore and
waste separation, since the 1960s, the development of high-gradient magnetic sepa-
rators (HGMS) has led to applications in a wide variety of industries. In the 1970s

Fig. 12 Magnetic pulley for


separation of strongly
Nonmagnetic
magnetic materials. (Adapted Material
from [222])
avel
Belt tr
nent
Perma
a g n e ts
M

Tramp iron
Adjustable
nonmagnetic
splitter
33 Magnetochemistry and Magnetic Separation 1617

and 1980s, research was primarily focused on water treatment [226], petroleum
and ore processing [227, 228], and magnetic separation for discrimination of air
pollutants [229]. While in the last three decades, magnetic separation has become of
great interest to wastewater treatment [230]. Traditionally, industrial separation has
been the source of the majority of interest, but since the 1970s [231], there has been a
realization that magnetic fields and magnetic carriers can be used in the life sciences
too [232], especially for the separation of cells and proteins [231, 233]. Thus
the application domains of magnetic separation can be categorized into a number
of resource management and bio-healthcare subdivisions. Resource management
primarily consists of industrial waste treatment and ore refinement [222, 234],
whereas bio-healthcare is concerned with clinical applications: biological matter
separation [231,233] and industrial separation and purification of pharmacologically
active drugs and industrially relevant bacteria. Within the resource management
domain, a number of general classifications for separator types have emerged over
the years, but the most commonly used one refers to the intensity of the magnetic
field gradients (high or low intensity), and to the separating medium (dry, wet),
along with a few other classifications (Table 5).

Table 5 Classification of magnetic separators [222]


Type Examples Applications
Dry low-intensity Magnetic pulleys, plate magnets, Primarily used to remove strongly
grate magnets, suspended magnets, magnetic tramp iron
drum magnets
Wet low-intensity Concurrent tank, counter-rotator These separators are used mainly
tank, counter-current tank for the recovery of heavy media,
such as magnetite or ferrosilicon
used in dense media separation
Dry high-intensity Cross-belt, induced magnetic roll, Applicable to the beneficiation of
permanent magnet roll coarse, weakly magnetic minerals
Wet high-intensity High gradient magnetic separators Used in a variety of industries, most
(HGMS) notably kaolin purification,
iron-ore, and beach sand
beneficiation
Superconducting Cyclic, superconducting HGMS, Higher throughput compared to
open gradient standard electromagnets
Eddy current Separates non-ferrous metals, eddy With this method, it is possible to
currents can be induced by applying sort different non-ferrous metals
alternating magnetic fields, which such as aluminium and copper
generates a repulsive force such that
the non-ferrous material is thrown
off the end of a conveyor belt
Magneto- Diamagnetic or weakly Separators of this type are used in
Archimedes paramagnetic material can act as diamond sorting
effect diamagnetic voids in highly
paramagnetic solutions that are
exposed to external fields (Fig. 9)
1618 P. Dunne

Ore refinement was one of the earliest applications of magnetic separation and
now includes the beneficiation and separation of iron and manganese, ores, ilmenite,
some non-ferrous ores, mineral sands, and coal [222,234]. Coal beneficiation (sulfur
removal) is one of the most crucial new areas of magnetic separation [235,236,237],
as rapidly developing countries such as China increase the global demand for coal.
Coal plays a predominant role in Chinese energy generation, accounting in 2011 for
77% [238, 239] of their primary energy production.
Recycling is the other major field of magnetic separation, which includes the
separation of scrap metal and electronic circuit boards from landfills [240, 241,
242]; recycling ferrous materials from slag, i.e., by-products from metallurgical
processes [243,244]; waste incineration, such as municipal waste [245]; and, finally
in the nuclear industry, the removal of solids from fuel dissolved liquids and soil
decontamination [228, 246, 234].

High-Gradient Magnetic Separation (HGMS)

One of the key developments in industrial magnetic separation was high- gradient
separation in the 1960s and 1970s [247]. A suspension of ferromagnetic or
paramagnetic material in liquid flowed through a fine soft ferromagnetic mesh or
steel wool in a conduit inside a large electromagnet, or superconducting magnet
(Fig. 13). The fine magnetic wires distort the local field, generating magnetic field
gradients of the order 105 T m−1 , and forces 1012 N m−3 , trapping the magnetic
particles so long as the external magnetic field is applied. This is followed by a
purge-clean cycle with the magnetic field off (Fig. 13b), or alternatively the filter
material can be replaced after heavy use.
HGMS is most effective for particles of diameter >10 nm [248], and to date, the
most successful industrial application of HGMS is the removal of magnetic impu-
rities from kaolin, or china clay, a naturally white aluminosilicate Al2 Si2 O5 (OH)4
mineral [228, 234]. Other popular applications include purification of wastewater
from factories [249,250], paper mills [251], and the removal of martensite stainless
steel particles from viscous liquids [252]. However, there are problems with HGMS
systems, the two most important being, first, an uncertainty in magnetic field
distributions, leading to difficulties in optimizing process conditions [250, 248],
and, secondly, reduced performance due to permanent aggregation of particles in
the filtration unit. The Fe2 O3 nanoparticles used for arsenic removal, for example,
can’t be removed from steel wires [253].

Magnetic Carriers

The success of HGMS and open-gradient separation has driven the development
of separation using magnetic carriers. Here the goal is to specifically adsorb the
material of interest onto magnetic nanoparticles (carriers) and then remove the
magnetic carriers via magnetophoresis [224]. This is shown in Fig. 14, where a
33 Magnetochemistry and Magnetic Separation 1619

a) Filtered stream b)
(F small)

Flush water off Flush water on

Canister Canister

Magnet on Magnet off

Steel Mesh

Feed stream Flush stream


(F large)
Feed stream
Off

Fig. 13 Cross-section schematic of a high gradient magnetic separation (HGSM) system and fluid
flow pattern in a cyclic HGMS system during (a) the filtration and (b) washing sequences

suspension of coated iron oxide nanoparticles are strongly attracted to the fringing
field of two NdFeB permanent magnets. This type of separation has a greater
specificity and higher efficiency compared to similar centrifugation or filtration
methods [254]. It is faster and has a lower energy consumption [230], reduced
costs, and less contamination produced [255, 256], and for reusable nanoparticles
acting as catalysts, separation is carried out under controlled environments, avoiding
exposure of the catalysts to air [230]. However, magnetic forces scale with volume,
so larger magnetic nanoparticles are desirable for permanent capture [234, 225];
with magnetite particles, the minimum size for permanent capture was 40 nm for
uncoated and 70 nm for coated nanoparticles [249].

Wastewater Treatment
In terms of wastewater treatment, most of the research emphasis has been on
the removal of heavy metals and organic dyes from solution. Heavy metal waste
products are produced in a large number of industrial processes such as metal ore
mining and refinement, metal finishing, painting, printing, and pesticide and fertil-
izer manufacturing [258, 259, 260]. They are toxic, carcinogenic, and mutagenic,
even at relatively low concentration [260, 261], and are non-biodegradable and
tend to accumulate in living tissues [262]. So their separation and purification from
1620 P. Dunne

Fig. 14 Magnetic separation of oleic acid (OA)-Pluronic-coated iron oxide magnetic nanoparti-
cles. (From [257])

wastewater is an ongoing problem attracting a lot of attention [263, 264]. Similarly,


organic dyes are toxic to aquatic life and can be both carcinogenic and mutagenic.
The issue is serious as there are more than 100,000 different dyes [230] in use in
a wide variety of industries including paper, plastic, leather, pharmaceutical, food,
cosmetic, and textile production [265, 266, 267, 268]. Many dyes have been banned
but are still used in some countries with weak legislation or supervision [269].
It is not economical to treat the waste as much of it is stable to light, heat, and
oxidizing agents in conventional water treatment plants [269]. These limitations
are particularly troublesome in the treatment of colored wastewater from textile
plants [270].
There have been some noted successes with nanoscale zero valence iron (nZVI),
and there are now pilot and full-scale industrial tests in Europe [271]. However,
there are still problems that need to be resolved before magnetic separation using
magnetic carriers is widely adopted in industry. This includes limited cyclability
of the nanoparticles, sometimes with as few as three cycles [230], meaning waste
management of the spent solutions and spent nanoparticles has to be considered. The
cost of production of these carriers on an industrial scale can vary from 225 $/kg for
iron oxides to 2255 $/kg for nZVI nanoparticles [272].
33 Magnetochemistry and Magnetic Separation 1621

Biological Separation
The other domain where magnetic carriers are of interest is in biology and medicine.
This includes separation of proteins [273] and DNA [274], pre-concentration of
biomaterials for assays [275], and cell separation [231]. This involves surface func-
tionalization to bind or capture specific targets [276], as outlined in Fig. 15. There
are already some biotech commercial applications involving protein purification and
cell sorting [277], and magnetic separation has been recently used in the processing
of bacteria for biofuel production [278]. See  Chap. 34, “Magnetism and Biology”
for further details.

Magnetically Stabilized Bed (MSB) Reactors


A fluidized bed reactor is one in which a gas or liquid is passed at high enough
velocities through a granular solid catalyst to suspend the solid and cause it to
behave like a fluid. As the gas or liquid velocity is increased beyond the minimum
fluidization velocity, most of the excess passes through the bed as bubbles, lowering
yields and leading to the entrainment of fine solids that need to be filtered out [163].
One way to improve their efficiency is to apply a magnetic field to a fluidized bed
reactor containing ferromagnetic or superparamagnetic particles parallel to the flow
velocity. Doing so can stabilize the fluidization process [280] by suppressing bubble
formation (Fig. 16), and combines the benefits of fixed bed (i.e., non-fluidized) and
conventional fluidized bed reactors by intensifying reaction processes, having a
smaller required pressure drop, high productivity, avoidance of particle clogging,
and minimal loss of solid particles [280, 281, 282]. There have also been reports
of a direct magnetic effect on catalyst efficiency, separate from the improved fluid
dynamics inside the reactor [283, 284, 285].
MSB reactors have been seen to improve a number of industrially important
processes including purification of caprolactam [286], Co2 reforming of methane
[287], immobilized enzyme reactions [282], fermentation of immobilized lipids
for biodiesel production [288, 289], hydrogenation of glucose into sorbitol [290],
hydrogenation of acetylene into ethane [291], chlorophenol dehalogenation [292],
olefin oligomerization [293], and CO methanation [294,295]. Since 2003, a number
of commercial MSB reactors are in operation such as a caprolactone hydrorefining
unit using amorphous Ni-based hydrogenation catalysts [290], and five industrial
MSB reactor units operating with a 2 Mt per year total production capacity, with six
more under construction in China [281].

Conclusions

The intersection of magnetism, (electro)chemistry, and fluid dynamics is a compli-


cated and sometimes chaotic area that continues to generate startling effects. Both
daughter disciplines of magnetoelectrochemistry and magnetic separation have long
and distinguished histories, yet both offer new and exciting challenges. In magneto-
electrochemistry, magnetic field gradient structuring of flows and deposits promises
a rich new world of lithography-free patterning and magnetically controlled and
1622

Fig. 15 A generalized procedure for magnetic separation for biological assays including carrier preparation. (From [279])
P. Dunne
33 Magnetochemistry and Magnetic Separation 1623

a b Stable Emulsion
Field Coil
Bubble

Solids
backmixing

Grid

Flow Flow

Fig. 16 Sketch comparing standard fluidized solids with magnetically stabilized ones in a
fluidized bed reactor. (a) The condition of a normal fluidized reactor at the minimum fluidization
velocity, containing a large number of voids and non-collinear local fluid velocities. (b) Magneti-
cally stabilized emulsion expands homogeneously to accommodate a higher throughflow. (Adapted
from [280])

stabilized microfluidics. Magnetohydrodynamic stirring during electrolysis allows


a one-time investment in a magnet system to improve production efficiency and
controllability for many deposition techniques, including electrowinning of Cu.
Whereas in magnetic separation, new developments in magnetic carriers have broad-
ened the horizon for cleaner, more efficient removal of heavy metals and dyes from
our waste-water, and ever more inventive ways of purifying drugs and separating
cells for our well-being. Both topics are converging to the same nanoscale region,
where the two most important questions are what is the smallest cohesive unit for
magnetophoresis and how can we exploit it? Indeed both magnetoelectrochemistry
and magnetic separation have a bright future, and no doubt there will be more
unexpected discoveries in the future.

Acknowledgments The author wishes to thank J.M.D. Coey for helpful comments and critiques
of this work.

References
1. Davidson, P.A.: An Introduction to Magnetohydrodynamics. Cambridge University Press,
Cambridge (2001)
2. Gunther, C.G.: Electro-Magnetic Ore Separation. Hill Publishing Company, New York (1909)
1624 P. Dunne

3. Coey, J.M.D.: Magnetism and Magnetic Materials. Cambridge University Press, Cambridge
(2010)
4. Burns, G.: Phys. Rev. 128(5), 2121 (1962)
5. Liebknecht, O., Wills, A.P.: Phys. Rev. (Series I) 10(4), 215 (1900)
6. March, H.W.: Phys. Rev. (Series I) 24(1), 29 (1907)
7. Brant, L.: Phys. Rev. 17(6), 678 (1921)
8. Woodbridge, D.B.: Phys. Rev. 48(8), 672 (1935)
9. Nettleton, H.R., Sugden, S.: Proc. R. Soc. Lond. A. Math. Phys. Sci. 173(954), 313 (1939)
10. Ikeda, T., Sueoka, K.: Bull. Chem. Soc. Jpn. 43, 1273 (1970)
11. Izutsu, K.: Electrochemistry in Nonaqueous Solutions. Wiley-VCH (2002)
12. Kittel, C.: Introduction to Solid State Physics, 7th edn. Wiley, New York (1996)
13. Selwood, P.W.: Magneochemistry, 1st edn. Interscience Publishers, New York (1943)
14. Bain, G.A., Berry, J.F.: J. Chem. Educ. 85(4), 532 (2008). https://doi.org/10.1021/ed085p532
15. Steiner, U.E., Ulrich, T.: Chem. Rev. 89(1), 51 (1989)
16. Hore, P.J., Mouritsen, H.: Annu. Rev. Biophys. 45(1), 299 (2016). https://doi.org/10.1146/
annurev-biophys-032116-094545
17. Zhang, Y., Berman, G.P., Kais, S.: Int. J. Quantum Chem. 115(19), 1327 (2015). https://doi.
org/10.1002/qua.24943
18. Lambert, N., Chen, Y.N., Cheng, Y.C., Li, C.M., Chen, G.Y., Nori, F.: Nat. Phys. 9(1), 10
(2013)
19. Rodgers, C.T., Hore, P.J.: Proc. Natl. Acad. Sci. U. S. A. 106(2), 353 (2009)
20. Messiha, H.L., Wongnate, T., Chaiyen, P., Jones, A.R., Scrutton, N.S.: J. R. Soc. Interface
12(103), 20141155 (2015)
21. Monti, D., Ottolina, G., Carrea, G., Riva, S.: Chem. Rev. 111(7), 4111 (2011)
22. Buetler, T.M.: News Physiol. Sci. 19(3), 120 (2004)
23. Shen, H.M., Liu, Z.G.: Free Radic. Biol. Med. 40(6), 928 (2006)
24. Taraban, M.B., Leshina, T.V., Anderson, M.A., Grissom, C.B.: J. Am. Chem. Soc. 119(24),
5768 (1997)
25. Jones, A.R., Scrutton, N.S., Woodward, J.R.: J. Am. Chem. Soc. 128(26), 8408 (2006)
26. Harkins, T.T., Grissom, C.B.: J. Am. Chem. Soc. 117(1), 566 (1995)
27. Jones, A.R., Hay, S., Woodward, J.R., Scrutton, N.S.: J. Am. Chem. Soc. 129(50), 15718
(2007)
28. Crotty, D., Silkstone, G., Poddar, S., Ranson, R., Prina-Mello, A., Wilson, M., Coey, J.M.D.:
Proc. Natl. Acad. Sci. 109(18), 7126 (2012)
29. Hore, P.J.: Proc. Natl. Acad. Sci. U. S. A. 109(5), 1357 (2012)
30. Maeda, K., Henbest, K.B., Cintolesi, F., Kuprov, I., Rodgers, C.T., Liddell, P.A., Gust, D.,
Timmel, C.R., Hore, P.J.: Nature 453(7193), 387 (2008)
31. Biskup, T., Schleicher, E., Okafuji, A., Link, G., Hitomi, K., Getzoff, E.D., Weber, S.: Angew.
Chemie – Int. Ed. 48(2), 404 (2009)
32. Hogben, H.J., Biskup, T., Hore, P.J.: Phys. Rev. Lett. 109, 1 (2012)
33. Schulten, K., Swenberg, C.E., Weller, A.: Zeitschrift für Phys. Chemie 111(1), 1 (1978)
34. Gross, W.: Verh. Dtsch. Phys. Ges. 38 (1885)
35. Hinds, G., Rhen, F.M.F., Coey, J.M.D.: IEEE Trans. Magn. 38(5), 3216 (2002)
36. Rhen, F.M.F., Hinds, G., Coey, J.M.D.: Electrochem. Commun. 6(4), 413 (2004)
37. Rhen, F.M.F., Fernandez, D., Hinds, G., Coey, J.M.D.: J. Electrochem. Soc. 153, J1 (2006)
38. Roberts, P.: How MHD transformed the theory of geomagnetism. In: Moreau, R. (ed.)
Magnetohydrodynamics. Fluid Mechanics and Its Applications, vol. 80, pp. 3–26. Springer,
Dordrecht (2007)
39. Yang, L.: J. Electrochem. Soc. 101, 456 (1954)
40. Aogaki, R., Fueki, K., Mukaibo, T.: Denki Kagaku oyobi Kogyo Butsuri Kagaku 43, 509
(1975)
41. Aogaki, R., Fueki, K., Mukaibo, T.: Denki Kagaku oyobi Kogyo Butsuri Kagaku 44, 89
(1976)
42. Bard, A.J., Faulkner, L.R.: Electrochemical Methods: Fundamentals and Applications, 2nd
edn. Wiley, New York/Chichester (2000)
33 Magnetochemistry and Magnetic Separation 1625

43. Marcus, R.A.: J. Chem. Phys. 24, 4966 (1956)


44. Hush, N.S.: J. Chem. Phys. 28, 962 (1958)
45. Levich, V.G.: Advances in electrochemistry and electrochemical engineering. In: Delahay, P.,
Tobias, C.W. (eds.) Advances in Electrochemistry and Electrochemical Engineering, vol. 4,
pp. 249–371. Wiley, New York (1966)
46. Dogonadze, R.R.: Theory of molecular electrode kinetics. In: Hush, N.H. (ed.) Reactions of
Molecules at Electrodes. Wiley-Interscience, New York (1971)
47. Kuznetsov, A.M.: Charge Transfer in Physics, Chemistry and Biology. Gordon & Breach
(2000)
48. Brett, C.M.A., Oliveira-Brett, A.M.: Electrochemistry: Principles, Methods, and Applica-
tions. Oxford University Press, Oxford/New York/Tokyo (2000)
49. Monzon, L.M., Coey, J.M.D.: Electrochem. Commun. 42, 42 (2014)
50. Coey, J.M.D., Aogaki, R., Byrne, F., Stamenov, P.: Proc. Natl. Acad. Sci. 106(22), 8811
(2009)
51. Aaboubi, O., Chopart, J.P., Douglade, J., Olivier, A., J. Electrochem. Soc. 137, 1796 (1990)
52. Qian, S., Bau, H.H.: Phys. Fluids 17(6), 067105 (2005)
53. Aogaki, R., Morimoto, R., Asanuma, M.: J. Magn. Magn. Mater. 322(9–12), 1664 (2010)
54. Gailitis, A., Lielausis, O., Platacis, E., Dement’ev, S., Cifersons, A., Gerbeth, G., Gundrum,
T., Stefani, F., Christen, M., Will, G.: Phys. Rev. Lett. 86, 3024 (2001)
55. Fahidy, T.Z.: J. Appl. Electrochem. 13(5), 553 (1983)
56. Tacken, R.A., Janssen, L.J.J.: J. Appl. Electrochem. 25(1), 1 (1995)
57. Fahidy, T.Z.: Prog. Surf. Sci. 68(4–6), 155 (2001)
58. Alemany, A., Chopart, J.P., Molokov, S., Moreau, R., Moffatt, K., Alemany, A., Chopart,
J.P.: An outline of magnetoelectrochemistry. In: Molokov, S., Moreau, R., Moffatt, K.
(eds.) Magnetohydrodynamics. Fluid Mechanics and Its Applications, vol. 80, pp. 391–407.
Springer, Dordrecht (2007)
59. Aogaki, R., Ito, M., Ogata, E.: Application of magnetic micro-fluid chip to chemical and
electrochemical reactions. In: PAMIR International Conference on Fundamental and Applied
MHD, Riga, pp. 65–72 (2005)
60. Monzon, L.M., Coey, J.M.D.: Electrochem. Commun. 42, 38 (2014)
61. Mogi, I., Kikuya, H., Watanabe, K., Awaji, S., Motokawa, M.: Chem. Lett. 25, 673 (1996)
62. Devos, O., Aaboubi, O., Chopart, J.P., Merienne, E., Olivier, A., Amblard, J.: J. Electrochem.
Soc. 145, 4135 (1998)
63. Coey, J.M.D., Hinds, G.: J. Alloys Compd. 326(1), 238 (2001)
64. Tabakovic, I., Riemer, S., Vas’ko, V., Sapozhnikov, V., Kief, M.: J. Electrochem. Soc. 150(9),
C635 (2003)
65. Uhlemann, M., Gebert, A., Herrich, M., Krause, A., Cziraki, A., Schultz, L.: Electrochim.
Acta 48(20–22), 3005 (2003)
66. Matsushima, H., Nohira, T., Mogi, I., Ito, Y.: Surf. Coatings Technol. 179(2–3), 245
(2004)
67. Msellak, K., Chopart, J.P., Jbara, O., Aaboubi, O., Amblard, J.: J. Magn. Magn. Mater. 281,
295 (2004)
68. Ní Mhíocháin, T.R., Coey, J.M.D.: Phys. Rev. E 69(6), 061404 (2004)
69. Bund, A., Kuehnlein, H.H.: J. Phys. Chem. B 109(42), 19845 (2005)
70. Krause, A., Hamann, C., Uhlemann, M., Gebert, A., Schultz, L.: J. Magn. Magn. Mater. 290–
291(Part 1), 261 (2005)
71. Motoyama, M., Fukunaka, Y., Kikuchi, S.: Electrochim. Acta 51(5), 897 (2005)
72. Tabakovic, I., Riemer, S., Sun, M., Vas’ko, V.A., Kief, M.T.: J. Electrochem. Soc. 152(12),
C851 (2005)
73. Uhlemann, M., Krause, A., Gebert, A.: J. Electroanal. Chem. 577(1), 19 (2005)
74. Matsushima, H., Bund, A., Plieth, W., Kikuchi, S., Fukunaka, Y.: Electrochim. Acta 53(1),
161 (2007)
75. Koza, J.A., Uhlemann, M., Gebert, A., Schultz, L.: Electrochim. Acta 53(16), 5344 (2008)
76. Matsushima, H., Ispas, A., Bund, A., Bozzini, B.: J. Electroanal. Chem. 615(2), 191 (2008)
77. Fernández, D., Coey, J.M.D.: Electrochem. Commun. 11(2), 379 (2009)
1626 P. Dunne

78. Koza, J.A., Karnbach, F., Uhlemann, M., McCord, J., Mickel, C., Gebert, A., Baunack, S.,
Schultz, L.: Electrochim. Acta 55(3), 819 (2010)
79. Chiba, A., Kitamura, K., Ogawa, T.: Surf. Coat. Technol. 27, 83 (1986)
80. Georgescu, V., Daub, M.: Surf. Sci. 600(18), 4195 (2006)
81. Krause, A., Uhlemann, M., Gebert, A., Schultz, L.: Thin Solid Films 515(4), 1694
(2006)
82. Chouchane, S., Levesque, A., Douglade, J., Rehamnia, R., Chopart, J.P.: Surf. Coat. Technol.
201(14), 6212 (2007)
83. Daltin, A.L., Bohr, F.F., Chopart, J.P.: Electrochim. Acta 54(24), 5813 (2009)
84. Lee, C.C., Chou, T.C.: Electrochim. Acta 40, 965 (1995)
85. Hinds, G., Coey, J.M.D., Lyons, M.E.G.: J. Appl. Phys. 83(11), 6447 (1998)
86. Sugiyama, A., Morisaki, S., Aogaki, R.: Jpn. J. Appl. Phys. 42(Part 1, No. 8), 5322 (2003)
87. Hinds, G., Spada, F.E., Coey, J.M.D., Ní Mhíocháin, T.R., Lyons, M.E.G.: J. Phys. Chem. B
105(39), 9487 (2001)
88. Hinds, G., Coey, J.M.D., Lyons, M.E.G.: Electrochem. Commun. 3(5), 215 (2001)
89. Coey, J.M.D.: Europhys. News 34(6), 246 (2003)
90. Fricoteaux, P., Jonvel, B., Chopart, J.P.: J. Phys. Chem. B 107(35), 9459 (2003)
91. Krause, A., Uhlemann, M., Gebert, A., Schultz, L.: Electrochim. Acta 49, 4127 (2004)
92. Legeai, S., Chatelut, M., Vittori, O., Chopart, J.P., Aaboubi, O.: Electrochim. Acta 50(1), 51
(2004)
93. Zhang, Y., Liu, M., Wang, M., Xie, Q., Yao, S.: Sensors Actuators B Chem. 123(1), 444
(2007)
94. Pané, S., Gómez, E., Vallés, E.: J. Electroanal. Chem. 615(2), 117 (2008)
95. Li, D., Levesque, A., Franczak, A., Wang, Q., He, J., Chopart, J.P.: Talanta 110, 66 (2013)
96. Żabiński, P.R., Franczak, A., Kowalik, R.: Arch. Metall. Mater. 57(2), 495 (2012)
97. Osaka, T., Homma, T., Saito, K., Takekoshi, A., Yamazaki, Y., Namikawa, T.: J. Electrochem.
Soc. 139, 1311 (1992)
98. Liu, C., Tian, A., Yang, H., Xu, Q., Xue, X.: Appl. Surf. Sci. 287, 218 (2013)
99. Rao, V.R., Bangera, K.V., Hegde, A.C.: J. Magn. Magn. Mater. 345, 48 (2013)
100. Żabiński, P., Mech, K., Kowalik, R.: Electrochim. Acta 104, 542 (2013)
101. Xu, Q., Liu, Y., Liu, C., Tian, A., Shi, X., Dong, C., Zhou, Y., Zhou, H.: RSC Adv. 5(19),
14458 (2015)
102. Koza, J.A., Uhlemann, M., Mickel, C., Gebert, A., Schultz, L.: J. Magn. Magn. Mater.
321(14), 2265 (2009)
103. Olivas, P., Alemany, A., Bark, F.H.: J. Appl. Electrochem. 34(1), 19 (2004)
104. Kabbani, H., Wang, A., Luo, X., Qian, S.: Phys. Fluids 19(8), 083604 (2007)
105. Weier, T., Cierpka, C., Hüller, J., Gerbeth, G.: Magnetohydrodynamics 42, 379 (2006)
106. Weier, T., Eckert, K., Mühlenhoff, S., Cierpka, C., Bund, A., Uhlemann, M.: Electrochem.
Commun. 9(10), 2479 (2007)
107. Mutschke, G., Bund, A.: Electrochem. Commun. 10(4), 597 (2008)
108. König, J., Mühlenhoff, S., Eckert, K., Büttner, L., Odenbach, S., Czarske, J.: Electrochim.
Acta 56(17), 6150 (2011)
109. König, J., Neumann, M., Mühlenhoff, S., Tschulik, K., Albrecht, T., Eckert, K., Uhlemann,
M., Weier, T., Büttner, L., Czarske, J.: Eur. Phys. J. Spec. Top. 220(1), 79 (2013)
110. Daltin, A.L., Chopart, J.P.: Cryst. Growth Des. 10(5), 2267 (2010)
111. Daltin, A.L., Addad, A., Baudart, P., Chopart, J.P.: Cryst. Eng. Commun. 13(10), 3373 (2011)
112. Koza, J.A., Mogi, I., Tschulik, K., Uhlemann, M., Mickel, C., Gebert, A., Schultz, L.:
Electrochim. Acta 55(22), 6533 (2010)
113. Koza, J.A., Uhlemann, M., Gebert, A., Schultz, L.: Electrochim. Acta 53(27), 7972 (2008)
114. Mühlenhoff, S., Mutschke, G., Koschichow, D., Yang, X., Bund, A., Fröhlich, J., Odenbach,
S., Eckert, K.: Electrochim. Acta 69, 209 (2012)
115. Krause, A., Koza, J., Ispas, A., Uhlemann, M., Gebert, A., Bund, A.: Electrochim. Acta
52(22), 6338 (2007)
33 Magnetochemistry and Magnetic Separation 1627

116. Kozuka, T., Sakurai, S., Kawahara, M.: Effect of magnetic field on surface morphology in
electrodeposition process. In: Proceedings of 6th International Conference on Electromagnetic
Processing of Materials, Dresden, pp. 431–434 (2009)
117. Asanuma, M., Yamada, A., Aogaki, R.: Jpn. J. Appl. Phys. 44(7A), 5137 (2005)
118. Aogaki, R., Miura, M., Oshikiri, Y.: ECS Trans. 16(25), 181 (2009)
119. Mogi, I., Morimoto, R., Aogaki, R., Watanabe, K.: Sci. Rep. 3, 2574 (2013)
120. Mogi, I., Aogaki, R., Morimoto, R., Watanabe, K.: ECS Trans. 45(12), 1 (2013)
121. Fernández, D., Diao, Z., Dunne, P., Coey, J.M.D.: Electrochim. Acta 55(28), 8664 (2010)
122. Koza, J.A., Uhlemann, M., Gebert, A., Schultz, L.: J. Electroanal. Chem. 617(2), 194
(2008)
123. Koza, J.A., Mühlenhoff, S., Uhlemann, M., Eckert, K., Gebert, A., Schultz, L.: Electrochem.
Commun. 11(2), 425 (2009)
124. Koza, J.A., Mühlenhoff, S., Żabiński, P., Nikrityuk, P., Eckert, K., Uhlemann, M., Gebert, A.,
Weier, T., Schultz, L., Odenbach, S.: Electrochim. Acta 56(6), 2665 (2011)
125. Fernández, D., Martine, M.L., Meagher, A., Möbius, M.E., Coey, J.M.D.: Electrochem.
Commun. 18, 28 (2012)
126. Diao, Z., Dunne, P.A., Zangari, G., Coey, J.M.D.: Electrochem. Commun. 11(4), 740
(2009)
127. Lin, M.Y., Hourng, L.W., Kuo, C.W.: Int. J. Hydrogen Energy 37(2), 1311 (2012)
128. Weier, T., Landgraf, S.: Eur. Phys. J. Spec. Top. 220(1), 313 (2013)
129. Matsushima, H., Iida, T., Fukunaka, Y.: Electrochim. Acta 100, 261 (2013)
130. Fernández, D., Maurer, P., Martine, M.L., Coey, J.M.D., Möbius, M.E.: Langmuir 30(43),
13065 (2014)
131. Liu, W., Cai, W., Liu, C., Sun, S., Xing, W.: Fuel 139, 308 (2015)
132. Leventis, N., Gao, X.: Anal. Chem. 73(16), 3981 (2001)
133. Kishioka, S.Y., Yamada, A., Aogaki, R.: Electroanalysis 13(14), 1161 (2001)
134. Lee, J., Ragsdale, S.R., Gao, X., White, H.S.: J. Electroanal. Chem. 422(1–2), 169 (1997)
135. Dunne, P., Coey, J.M.D.: 123(39), 24181 (2019). https://doi.org/10.1021/acs.jpcc.9b07534
136. Waskaas, M.: J. Phys. Chem. 97, 6470 (1993)
137. Waskaas, M.: Acta Chem. Scan. 50(6), 516 (1996)
138. Waskaas, M., Kharkats, Y.I.: J. Phys. Chem. B 103(23), 4876 (1999)
139. Perov, N.S., Sheverdyaeva, P.M., Inoue, M.: J. Appl. Phys. 91(10), 8557 (2002)
140. Perov, N.S., Sheverdyaeva, P.M., Inoue, M.: J. Magn. Magn. Mater. 272–276(Part 3), 2448
(2004)
141. Rhen, F.M.F., Coey, J.M.D.: J. Phys. Chem. B 110(12), 6274 (2006)
142. Rhen, F.M.F., Coey, J.M.D.: J. Phys. Chem. C 111(8), 3412 (2007)
143. Lu, Z.: ECS Trans. 59(1), 429 (2014)
144. Shinohara, K., Aogaki, R.: Electrochemistry 67, 126 (1999)
145. Shinohara, K., Hashimoto, K., Aogaki, R.: Chem. Lett. 31, 778 (2002)
146. Asanuma, M., Yamada, A., Aogaki, R.: Jpn. J. Appl. Phys. 43(9A), 6303 (2004)
147. Sueptitz, R., Tschulik, K., Uhlemann, M., Gebert, A., Schultz, L.: Electrochim. Acta 55(18),
5200 (2010)
148. Sueptitz, R., Tschulik, K., Uhlemann, M., Schultz, L., Gebert, A.: Corros. Sci. 53(10), 3222
(2011)
149. Sueptitz, R., Tschulik, K., Uhlemann, M., Schultz, L., Gebert, A.: Electrochim. Acta 56(17),
5866 (2011)
150. Monzon, L.M.A., Rode, K., Venkatesan, M., Coey, J.M.D.: Chem. Mater. 24(20), 3878 (2012)
151. Liu, M.: Phys. Rev. Lett. 84(12), 2762 (2000)
152. Landau, L.D., Pitaevskii, L.P., Lifshitz, E.M.: Electrodynamics of Continuous Media, 2nd
edn. Elsevier Butterworth-Heinemann, Oxford (2004)
153. Coey, J.M.D., Rode, K.: Dilute magnetic oxides and nitrides. In: Handbook of Magnetism
and Advanced Magnetic Materials. Wiley, Chichester (2007)
154. O’Brien, R.N., Santhanam, K.S.V.: J. Appl. Electrochem. 27, 573 (1997)
1628 P. Dunne

155. Yang, Y., Grant, K.M., White, H.S., Chen, S.: Langmuir 19(22), 9446 (2003)
156. Rabah, K.L., Chopart, J.P., Schloerb, H., Saulnier, S., Aaboubi, O., Uhlemann, M., Elmi, D.,
Amblard, J.: J. Electroanal. Chem. 571(1), 85 (2004)
157. Bund, A., Ispas, A.: J. Electroanal. Chem. 575(2), 221 (2005)
158. Sugiyama, A., Hashiride, M., Morimoto, R., Nagai, Y., Aogaki, R.: Electrochim. Acta 49(28),
5115 (2004)
159. Coey, J.M.D., Rhen, F.M.F., Dunne, P., McMurry, S.: J. Solid State Electrochem. 11(6), 711
(2007)
160. Mutschke, G., Cierpka, C., Weier, T., Eckert, K., Mühlenhoff, S., Bund, A.: On three-
dimensional magnetic field effects during metal deposition in cuboid cells. In: ECS
Transactions, vol. 13, pp. 9–13. ECS (2008)
161. Mühlenhoff, S., Mutschke, G., Uhlemann, M., Yang, X., Odenbach, S., Fröhlich, J., Eckert,
K.: Electrochem. Commun. 36, 80 (2013)
162. Yang, X., Mühlenhoff, S., Nikrityuk, P.A., Eckert, K.: Eur. Phys. J. Spec. Top. 220(1), 303
(2013)
163. Rosensweig, R.: Nature 210(5036), 613 (1966)
164. Zimmels, Y., Tuval, Y., Lin, I.: IEE Trans. Magn. 13(4), 1045 (1977)
165. Dunne, P.A., Hilton, J., Coey, J.M.D.: J. Magn. Magn. Mater. 316(2), 273 (2007)
166. Catherall, A.T., López-Alcaraz, P., Benedict, K.A., King, P.J., Eaves, L.: New J. Phys. 7(1),
118 (2005)
167. Beaugnon, E., Tournier, R.: Nature 349(7), 470 (1991)
168. Simon, A.G.M., Boamfa, M., Heflinge, L.: Nature 400(6742), 323 (1999)
169. Beaugnon, E., Fabregue, D., Billy, D., Nappa, J., Tournier, R.: Phys. B 294–295, 715
(2001)
170. Motokawa, M., Hamai, M., Sato, T., Mogi, I., Awaji, S., Watanabe, K., Kitamura, N.,
Makihara, M.: Phys. B 294, 729 (2001)
171. Motokawa, M., Hamai, M., Sato, T., Mog, I., Awaj, S., Watanabe, K., Kitamura, N., Makihara,
M.: J. Mag. Mat. Mat. 226–230, 2090 (2001)
172. Chun, Y.D., Lee, J.: Phys. C 372–376, 1491 (2002)
173. Hirota, N., Ikezoe, Y., Uetake, H., Kaihatsu, T., Takayama, T., Kitazawa, K.: RIKEN Rev. 44,
159 (2002)
174. Mutschke, G., Tschulik, K., Weier, T., Uhlemann, M., Bund, A., Fröhlich, J.: Electrochim.
Acta 55(28), 9060 (2010)
175. Dunne, P., Mazza, L., Coey, J.M.D.: Phys. Rev. Lett. 107(2), 024501 (2011)
176. Ragsdale, S.R., Grant, K.M., White, H.S.: J. Am. Chem. Soc. 120, 13461 (1998)
177. Grant, K.M., Hemmert, J.W., White, H.S.: Electrochem. Commun. 1(8), 319 (1999)
178. Pullins, M.D., Grant, K.M., White, H.S.: J. Phys. Chem. B 105(37), 8989 (2001)
179. Grant, K.M., Hemmert, J.W., White, H.S.: J. Am. Chem. Soc. 124(3), 462 (2002)
180. Mehta, D., White, H.S.: Chem. Eur. J. Chem. Phys. 4(2), 212 (2003)
181. Aogaki, R., Ito, E., Ogata, M.: J. Solid State Electrochem. 11(6), 757 (2007)
182. Deng, T., Prentiss, M., Whitesides, G.M.: Appl. Phys. Lett. 80(3), 461 (2002)
183. Inglis, D.W., Riehn, R., Sturm, J.C., Austin, R.H.: J. Appl. Phys. 99(8), 08K101 (2006)
184. Bau, H.H., Zhu, J., Qian, S., Xiang, Y.: Sensors Actuators B Chem. 88(2), 205 (2003)
185. Martins, A.A., Pinheiro, M.J.: Phys. Fluids 21(9), 097103 (2009)
186. Qian, S., Bau, H.H.: Mech. Res. Commun. 36(1), 10 (2009)
187. Anderson, E.C., Weston, M.C., Fritsch, I.: Anal. Chem. 82(7), 2643 (2010)
188. Weston, M.C., Gerner, M.D., Fritsch, I.: Anal. Chem. 82(9), 3411 (2010)
189. Weston, M.C., Nash, C.K., Fritsch, I.: Anal. Chem. 82(17), 7068 (2010)
190. Fritsch, V.S.I.: Microfluid. Nanofluidics 18, 159 (2015)
191. Sahore, V., Fritsch, I.: Microfluid. Nanofluidics 18(2), 159 (2014)
192. Gorobets, O.Y., Gorobets, V.Y., Derecha, D.O., Brukva, O.M.: J. Phys. Chem. C 112(9), 3373
(2008)
193. Tschulik, K., Sueptitz, R., Koza, J., Uhlemann, M., Mutschke, G., Weier, T., Gebert, A.,
Schultz, L.: Electrochim. Acta 56(1), 297 (2010)
33 Magnetochemistry and Magnetic Separation 1629

194. Tschulik, K., Yang, X., Mutschke, G., Uhlemann, M., Eckert, K., Sueptitz, R., Schultz, L.,
Gebert, A.: Electrochem. Commun. 13(9), 946 (2011)
195. Dunne, P., Coey, J.M.D.: Phys. Rev. B 85(22), 224411 (2012)
196. Tschulik, K., Sueptitz, R., Uhlemann, M., Schultz, L., Gebert, A.: Electrochim. Acta 56(14),
5174 (2011)
197. Dunne, P., Coey, J.M.D.: Phys. Rev. Lett. 109(22), 229402 (2012)
198. Mutschke, G., Tschulik, K., Uhlemann, M., Bund, A., Fröhlich, J.: Phys. Rev. Lett. 109(22),
229401 (2012)
199. Fujiwara, M., Chie, K., Sawai, J., Shimizu, D., Tanimoto, Y.: J. Phys. Chem. B 108(11), 3531
(2004)
200. Uhlemann, M., Tschulik, K., Gebert, A., Mutschke, G., Fröhlich, J., Bund, A., Yang, X.,
Eckert, K.: Eur. Phys. J. Spec. Top. 220(1), 287 (2013)
201. Yang, X., Tschulik, K., Uhlemann, M., Odenbach, S., Eckert, K.: IEEE Trans. Magn. 50(11),
1 (2014)
202. Pulko, B., Yang, X., Lei, Z., Odenbach, S., Eckert, K.: Appl. Phys. Lett. 105(23), 232407
(2014)
203. Eliassen, R., Skrinde, R.T., Davis, W.B., Eliassen, R., Skrinde, R.T., Davis, W.B.: J. Am.
Water Work. Assoc. 50, 1371 (1958)
204. Kronenberg, K.: Magn. IEEE Trans. 21(5), 2059 (1985)
205. Cai, R., Yang, H., He, J., Zhu, W.: J. Mol. Struct. 938(1–3), 15 (2009)
206. Amiri, M.C., Dadkhah, A.A.: Colloids Surfaces A Physicochem. Eng. Asp. 278(1–3), 252
(2006)
207. Toledo, E.J., Ramalho, T.C., Magriotis, Z.M.: J. Mol. Struct. 888(1–3), 409 (2008)
208. Lee, S.K.S.H., Jeon, S.I., Kim, Y.S., Lee, S.K.S.H.: J. Mol. Liq. 187, 230 (2013)
209. Pang, X.F., Deng, B.: Phys. B Condens. Matter 403(19–20), 3571 (2008)
210. Hołysz, L., Chibowski, M., Chibowski, E.: Colloids Surfaces A Physicochem. Eng. Asp.
208(1–3), 231 (2002)
211. Silva, I.B., Queiroz Neto, J.C., Petri, D.F.: Colloids Surfaces A Physicochem. Eng. Asp. 465,
175 (2015)
212. Higashitani, K., Oshitani, J.: J. Colloid Interface Sci. 204(2), 363 (1998)
213. Coey, J.M.D., Cass, S.: J. Magn. Magn. Mater. 209(1–3), 71 (2000)
214. Szcześ, A., Chibowski, E., Hołysz, L., Rafalski, P.: J. Phys. Chem. A 115(21), 5449 (2011)
215. Barrett, R.A., Parsons, S.A.: Water Res. 32(3), 609 (1998)
216. Powell, M.J.: Skept. Inq. 22(1), 27 (1998)
217. Demichelis, R., Raiteri, P., Gale, J.D., Quigley, D., Gebauer, D.: Nature Commun. 2, 590
(2011)
218. Gebauer, D., Völkel, A., Cölfen, H.: Science 322(5909), 1819 (2008)
219. Coey, J.M.D.: Philos. Mag. 92(31), 3857 (2012)
220. Zhou, K.X., Lu, G.W., Zhou, Q.C., Song, J.H., Jiang, S.T., Xia, H.R.: J. Appl. Phys. 88(2000),
1802 (2000)
221. Chang, K.T., Weng, C.I.: J. Appl. Phys. 100, 043917 (2006)
222. Svoboda, J., Fujita, T.: Miner. Eng. 16(9), 785 (2003)
223. Khashan, S.A., Furlani, E.P.: Microfluid. Nanofluidics 12(1–4), 565 (2012)
224. Friedman, G., Yellen, B.: Curr. Opin. Colloid Interface Sci. 10(3–4), 158 (2005)
225. Faraudo, J., Andreu, J.S., Camacho, J.: Soft Matter 9(29), 6654 (2013)
226. De Latour, C.: IEEE Trans. Magn. 9(3), 314 (1973)
227. Oberteuffer, J.: IEEE Trans. Magn. 10(2), 223 (1974)
228. Parker, M.: Phys. Technol. 12(6), 263 (1981)
229. Oldfield, F., Hunt, A., Jones, M.D.H., Chester, R., Dearing, J.A., Olsson, L., Prospero, J.M.:
Nature 317(6037), 516 (1985)
230. Gómez-Pastora, J., Bringas, E., Ortiz, I.: Chem. Eng. J. 256, 187 (2014)
231. Melville, D., PAUL, F., Roath, S.: Nature 255(5511), 706 (1975)
232. Yiacoumi, S., Rountree, D.A., Tsouris, C.: J. Colloid Interface Sci. 184(2), 477 (1996)
233. Molday, R.S., Yen, S.P.S., Rembaum, A.: Nature 268(5619), 437 (1977)
1630 P. Dunne

234. Yavuz, C.T., Prakash, A., Mayo, J., Colvin, V.L.: Chem. Eng. Sci. 64(10), 2510 (2009)
235. Zhang, Q., Gui, K.: J. Hazard. Mater. 168(2–3), 1341 (2009)
236. Luo, Z., Zhao, Y., Chen, Q., Tao, X., Fan, M.: Fuel Process. Technol. 85(2–3), 173 (2004)
237. Zhang, B., Zhao, Y., Duan, C., Tang, L., Dong, L., Qu, J.: Adv. Powder Technol. 25(3), 1031
(2014)
238. China NBS: Statistical Communiqué of the People’s Republic of China on the 2013 National
Economic and Social Development. Technical report, National Bureau of Statistics of China,
Beijing (2014)
239. IEA: Key World Energy Statistics 2014. OECD Publishing, Paris (2014)
240. Oliveira, C.R.D., Bernardes, A.M., Gerbase, A.E., de Oliveira, C.R., Bernardes, A.M.,
Gerbase, A.E.: Waste Manag. 32(8), 1592 (2012)
241. Veit, H.M., Diehl, T.R., Salami, A.P., Rodrigues, J.S., Bernardes, A.M., Tenório, J.A.S.: Waste
Manag. 25, 67 (2005)
242. Cui, J., Forssberg, E.: J. Hazard. Mater. 99(3), 243 (2003)
243. Fregeau, E.-W., Pignolet-Brandom, I., Iwasaki, S.: Liberation analysis of slow-cooled
steelmaking slags: implications for phosphorus removal. In: Proceedings of 1st International
Conference on Processing of Materology and Properties, pp. 153–156 (1993)
244. Das, B., Mohanty, J.K., Reddy, P.S.R., Ansari, M.I.: Scand. J. Metall. 26(4), 153 (1997)
245. Schmelzer, G.: Proc. XIX Int. Miner. Process. Congr. 4, 137 (1995)
246. Rossi, L.M., Garcia, M.A.S., Vono, L.L.R.: J. Braz. Chem. Soc. 23(11), 1959 (2012)
247. Oberteuffer, J.A., Wechsler, I., Marston, P.G., McNallan, M.J.: IEE Trans. Magn. 11(5), 1591
(1975)
248. Andreu, J.S., Camacho, J., Faraudo, J., Benelmekki, M., Rebollo, C., Martínez, L.M.: Phys.
Rev. E 84(2), 021402 (2011)
249. Moeser, G.D., Roach, K.A., Green, W.H., Alan Hatton, T., Laibinis, P.E.: AIChE J. 50(11),
2835 (2004)
250. Mariani, G., Fabbri, M., Negrini, F., Ribani, P.L.: Sep. Purif. Technol. 72(2), 147 (2010)
251. Kakihara, Y., Fukunishi, T., Takeda, S., Nishijima, S., Nakahira, A.: IEEE Trans. Appl.
Supercond. 14(2), 1565 (2004)
252. Mishima, F., Takeda, S., Fukushima, M., Nishijima, S.: Phys. C Supercond. 463–465, 1302
(2007)
253. Mayo, J., Yavuz, C., Yean, S., Cong, L., Shipley, H., Yu, W., Falkner, J., Kan, A., Tomson,
M., Colvin, V.: Sci. Technol. Adv. Mater. 8(1–2), 71 (2007)
254. Yavuz, C.T., Mayo, J.T., Yu, W.W., Prakash, A., Falkner, J.C., Yean, S., Cong, L., Shipley,
H.J., Kan, A., Tomson, M., Natelson, D., Colvin, V.L.: Science (80-). 314, 964 (2006)
255. Nassar, N.N.: Sep. Sci. Technol. 45(8), 1092 (2010)
256. Tan, Y., Chen, M., Hao, Y.: Chem. Eng. J. 191, 104 (2012)
257. Jain, T.K., Reddy, M.K., Morales, M.A., Leslie-Pelecky, D.L., Labhasetwar, V.: Mol. Pharm.
5(2), 316 (2008)
258. Kyzas, G.Z., Deliyanni, E.A.: Molecules 18(6), 6193 (2013)
259. Hao, Y.M., Man, C., Hu, Z.B.: J. Hazard. Mater. 184(1–3), 392 (2010)
260. Ahmed, M., Ali, S.M., El-Dek, S.I., Galal, A.: Mater. Sci. Eng. B 178(10), 744 (2013)
261. Bringas, E., San Román, M.F., Ortiz, I.: Ind. Eng. Chem. Res. 45(12), 4295 (2006)
262. Badruddoza, A.Z.M., Shawon, Z.B.Z., Tay, W.J.D., Hidajat, K., Uddin, M.S.: Carbohydr.
Polym. 91(1), 322 (2013)
263. Tang, S.C.N., Yan, D.Y.S., Lo, I.M.C.: Ind. Eng. Chem. Res. 53(40), 15718 (2014)
264. Singh, S., Barick, K.C., Bahadur, D.: J. Hazard. Mater. 192(3), 1539 (2011)
265. Liu, R., Shen, X., Yang, X., Wang, Q., Yang, F.: J. Nanoparticle Res. 15(6), 1679 (2013)
266. Ge, F., Ye, H., Li, M.M., Zhao, B.X.: Chem. Eng. J. 198–199, 11 (2012)
267. Debrassi, A., Correa, A.F., Baccarin, T., Nedelko, N., Slawska-Waniewska, A., Sobczak, K.,
Dluzewski, P., Greneche, J.M., Rodrigues, C.A.: Chem. Eng. J. 183, 284 (2012)
268. Anglada, A., Rivero, M.J., Ortíz, I., Urtiaga, A.: J. Chem. Technol. Biotechnol. 83(10), 1339
(2008)
269. Afkhami, A., Moosavi, R.: J. Hazard. Mater. 174(1–3), 398 (2010)
33 Magnetochemistry and Magnetic Separation 1631

270. Iram, M., Guo, C., Guan, Y., Ishfaq, A., Liu, H.: J. Hazard. Mater. 181(1–3), 1039 (2010)
271. Mueller, N.C., Braun, J., Bruns, J., Černík, M., Rissing, P., Rickerby, D., Nowack, B.: Environ.
Sci. Pollut. Res. Int. 19(2), 550 (2012)
272. Chowdhury, S.R., Yanful, E.K.: J. Environ. Manage. 91(11), 2238 (2010)
273. Becker, J.S., Thomas, O.R.T., Franzreb, M.: Sep. Purif. Technol. 65(1), 46 (2009)
274. Xie, X., Nie, X., Yu, B., Zhang, X.: J. Magn. Magn. Mater. 311(1), 416 (2007)
275. Safarik, I., Ptackova, L., Safarikova, M.: Eur. Cells Mater. 3(2), 52 (2002)
276. Berry, C.C., Curtis, A.S.G.: J. Phys. D. Appl. Phys. 36(13), R198 (2003)
277. Corchero, J.L., Villaverde, A.: Trends Biotechnol. 27(8), 468 (2009)
278. Lim, J.K., Chieh, D.C.J., Jalak, S.A., Toh, P.Y., Yasin, N.H.M., Ng, B.W., Ahmad, A.L.: Small
8(11), 1683 (2012)
279. He, J., Huang, M., Wang, D., Zhang, Z., Li, G.: J. Pharm. Biomed. Anal. 101, 84 (2014)
280. Rosensweig, R.E.: Science (80-.). 204(4388), 57 (1979)
281. Zong, B., Meng, X., Mu, X., Zhang, X.: Chinese J. Catal. 34(1), 61 (2013)
282. Bahar, T., Çelebi, S.S.: Enzyme Microb. Technol. 26(1), 28 (2000)
283. Lin, P.Y., Peng, J., Hou, B.H., Fu, Y.L.: J. Phys. Chem. 97(8), 1471 (1993)
284. Yermakov, A., Feduschak, T., Uimin, M., Mysik, A., Gaviko, V., Chupakhin, O., Shishmakov,
A., Kharchuk, V., Petrov, L., Kotov, Y., Vosmerikov, A., Korolyov, A.: Solid State Ionics
172(1–4), 317 (2004)
285. Sá, J., Szlachetko, J., Sikora, M., Kavčič, M., Safonova, O.V., Nachtegaal, M.: Nanoscale
5(18), 8462 (2013)
286. Meng, X., Mu, X., Zong, B., Min, E., Zhu, Z., Fu, S., Luo, Y.: Catal. Today 79–80, 21 (2003)
287. Hao, Z., Zhu, Q., Jiang, Z., Li, H.: Powder Technol. 183(1), 46 (2008)
288. Zhou, G.X., Chen, G.Y., Yan, B.B.: Biotechnol. Lett. 36(1), 63 (2014)
289. Liu, C.Z., Wang, F., Ou-Yang, F.: Bioresour. Technol. 100(2), 878 (2009)
290. Zong, B.: Catal. Surv. Asia 11(1–2), 87 (2007)
291. Dong, M., Pan, Z., Peng, Y., Meng, X., Mu, X., Zong, B., Zhang, J.: AIChE J. 54(5), 1358
(2008)
292. Graham, L.J., Atwater, J.E., Jovanovic, G.N.: AIChE J. 52(3), 1083 (2006)
293. Peng, Y., Dong, M., Meng, X., Zong, B., Zhang, J.: AIChE J. 55(3), 717 (2009)
294. Pan, Z., Dong, M., Meng, X., Zhang, X., Mu, X., Zong, B.: Chem. Eng. Sci. 62(10), 2712
(2007)
295. Li, J., Zhou, L., Zhu, Q., Li, H.: Ind. Eng. Chem. Res. 52(20), 6647 (2013)

Peter Dunne received his PhD from Trinity College Dublin in


2011. He worked at the Leibniz Institute for Solid State and
Materials Research in Dresden, Swiss Federal Laboratories for
Materials Science and Technology at Thun, and the Institute
of Supramolecular Science and Engineering in Strasbourg. He
now works on magneto-microfluidics and coupling spintronics to
interfacial chemistry.
Magnetism and Biology
34
Nora M. Dempsey

Contents
Biomagnetism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1634
Magnetic Navigation in Living Organisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1634
Magnetic Fields Produced by the Body . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1639
Iron in the Body . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1642
Magnetobiology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1643
Magnetic Nanoparticles for Biomedical Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1644
Magnetic Imaging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1646
Magnetic Fixation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1650
Magnetic Actuation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1652
Magnetic Heating . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1664
Effects of Magnetic Fields on Biological Organisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1665
Transcranial Magnetic Stimulation (TMS) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1667
Recommended Magnetic Field Exposure Limits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1667
Conclusions and Prospects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1670
Further Reading . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1670
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1671

Abstract

There are many links between magnetism and biology. One way to classify
them is according to whether they are naturally occurring or whether they
concern the exploitation of magnetism in biology and medicine. Here we refer
to the former as “biomagnetism” and the latter as “magnetobiology.” Examples
of biomagnetism discussed here are magnetic navigation in animals, magnetic
signals generated by animals, and the presence of iron in the body. The field of
magnetobiology may be subdivided into imaging, fixation, actuation, heating, or

N. M. Dempsey ()
Institut Néel, CNRS & Université Grenoble Alpes, Grenoble, France
e-mail: nora.dempsey@neel.cnrs.fr

© Springer Nature Switzerland AG 2021 1633


J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6_36
1634 N. M. Dempsey

stimulation, and it covers topics ranging from fundamental studies in biology


to biomedical applications. Biomagnetic phenomena are not new, but recent
advances in magnetobiology are shedding light on our current understanding of
biomagnetism. Thus, biomagnetism and magnetobiology are closely interlinked.
While the fundamental processes at play are well understood in certain applica-
tions (e.g., magnetic resonance imaging, cell sorting in microfluidics), the exact
link between the magnetic cause and the biological effect remains to be verified in
other cases (e.g., magnetically guided navigation in birds, transcranial magnetic
stimulation, etc.).

Biomagnetism

The study of naturally occurring magnetic phenomena in humans and animals has
a checkered past. Mesmer’s claim to harness a biophysical force he termed “animal
magnetism” to heal the so-called nervous diseases was discredited by a Royal
Commission in 1784. Nevertheless, many practitioners of alternative medicine still
claim to exploit biomagnetic energy in the treatment of a range of conditions.
Here we discuss three fields of study of naturally occurring magnetic phenomena
which have a solid scientific basis and which benefit from recent and ongoing
developments in magnetic and structural characterization tools and protocols.

Magnetic Navigation in Living Organisms

A number of living organisms, ranging from magnetotactic bacteria to insects


(honey bees, monarch butterflies), crustaceans (spiny lobsters), and vertebrates
including birds, amphibians, fish, and turtles, are known to use the Earth’s magnetic
field to help them navigate. The geomagnetic field is a vector field, and at any
particular spot on the Earth, it is characterized by its intensity (the average value
is of the order of 50 μT) and direction, including the inclination angle, which
is maximum near the magnetic poles and zero at the magnetic equator. In the
magnetic compass of birds, only the field’s axis (irrespective of polarity) is relevant
for navigation; in other animals, the field polarity (north or south) plays a role
and in certain cases the field intensity is pertinent. Our knowledge of how living
organisms are affected by the field is based on behavioral studies, in which scientists
observe the action of a given species in the local geomagnetic field and in an
artificially modified field, typically produced using Helmholtz coils. Further insight
is then gained by studying the physics, chemistry, and biology at play. Most
studies by far have been carried out on magnetotactic bacteria and migratory birds.
While magnetic orientation of magnetotactic bacteria is well understood, there is
still debate about the mechanism responsible for magnetic navigation in birds, as
detailed below.
34 Magnetism and Biology 1635

Magnetotactic Bacteria
Certain bacteria, referred to as magnetotactic bacteria [1, 2], move along the
geomagnetic field lines. This “magnetotaxis” was discovered when bacteria on
a microscope slide in Italy were observed to swim northward along the Earth’s
magnetic field lines [3]. Many types of magnetotactic bacteria have been found
throughout the world, in both fresh- and saltwater environments. They have been
shown to contain intracellular structures known as magnetosomes, which consist
of iron-containing crystals encapsulated in a membrane. The crystals, which are
synthesized by the bacteria, typically form intracellular chains (Fig. 1), though
clusters have also been observed. The vast majority of freshwater magnetotactic
bacteria contain nanocrystals of magnetite (Fe3 O4 ), while marine, estuarine, and
salt marsh species have been shown to contain the ferrimagnetic sulfospinel greigite
(Fe3 S4 ). The size of the crystals typically varies from 35 nm to 120 nm, which is
within the single-domain particle size range. Thus, magnetotactic bacteria naturally
produce permanently magnetized linear arrays of fully magnetized crystalites, and
it is the dipolar interaction that causes neighboring crystals in a chain to align
their moments along the chain axis, so as to minimize the magnetostatic energy
and maximize the magnetic moment of the organism. Coercivities of the order of
20–50 mT have been measured on cultured strains containing magnetite crystals.
The torque produced by the geomagnetic field on a magnetosome serves to align
the bacterial cell body parallel to the field, while movement along the field line is
produced by rotation of propeller-like helical flagella protruding from the bacteria.
The effect of the field is thus passive, and dead magnetotactic bacteria align with
the field but don’t move along it. The direction of motion of live bacteria is
switched when the direction of rotation of the flagella changes from clockwise
to counterclockwise, and vice versa. There is a predominance of north-seeking

Fig. 1 TEM image of a


magnetotactic bacterial cell
(Magnetospirillum) [5]. (With
kind permission of Springer
Nature)
1636 N. M. Dempsey

magnetotactic bacteria in the northern hemisphere, where the geomagnetic field


lines are inclined downward, and south-seeking magnetotactic bacteria in the south-
ern hemisphere, where the field lines are inclined upward. It was originally thought
that magnetotaxis simply served to guide magnetotactic bacteria down toward
oxygen-poor anaerobic sediments and thus away from surface water containing
toxic levels of oxygen. However, this does not explain the documented existence of
south-seeking magnetotactic bacteria in the northern hemisphere nor the discovery
of large populations of magnetotactic bacteria in the oxic–anoxic transition zone,
corresponding to the transition region between oxygen-rich and oxygen-poor water
or sediment. Aerotaxis refers to the movement of bacteria to an optimal oxygen
concentration in an oxygen gradient. Laboratory-based experiments involving
oxygen gradients have been used to show that in the bacteria, magnetotaxis works
together with aerotaxis. Two different magneto-aerotactic mechanisms have been
identified in different species, polar magneto-aerotaxis and axial magneto-aerotaxis.
In the former case, the geomagnetic field provides an axis and direction for motility,
while in the latter, it provides only an axis. The large majority of species display
polar magnetotaxis. In all cases, magnetotaxis serves to reduce a three-dimensional
search problem to a one-dimensional search along the geomagnetic field lines and
thus increases bacteria’s efficiency in finding their preferred habitat in a chemically
stratified environment such as sediment. When magnetotactic bacteria die and their
magnetosomes become trapped in sediments, the magnetite crystals may become
fossilized. These magnetofossils contribute to the natural remanent magnetization of
sediment layers and can be used as an indicator of the geomagnetic field orientation
at the time the sediment consolidated, provided the magnetofossils survive reductive
dissolution. Such sedimentary magnetofossils bear witness to how the Earth’s
magnetic polarity has switched, an event that occurs irregularly, but on average every
450,000 years. Magnetofossils are also studied as paleoindicators of past variations
of environmental temperature and humidity, which are believed to have affected
production of the bacteria. The relatively high structural perfection and narrow size
range of crystals have even prompted proposals to use magnetotactic bacteria as
bio-factories to produce magnetic nanoparticles [4].

Avian Magnetic Navigation


Migratory birds are known to use the geomagnetic field to navigate between
their summer and winter habitats [6, 7]. Behavioral studies in which the vertical
component of the magnetic field felt by caged European robins was inverted
demonstrated that the robins’ magnetic compass is sensitive to the axial orientation
of the magnetic field lines but not their polarity [8]. Thus, it is an inclination compass
that appears to measure the angle between field lines and the gravity vector to
indicate polarward or equatorward. It has also been shown that the avian magnetic
compass is light-dependent, with a given species able to navigate correctly when
exposed to light of a certain wavelength (from UV to green in the case of European
robins), while they become disoriented when exposed to light of a longer wavelength
[9]. Two mechanisms have been suggested to explain magnetoreception in birds.
34 Magnetism and Biology 1637

Magnetite-Based Magnetoreception
In analogy to magnetotactic bacteria, the magnetite hypothesis assumes a special-
ized sensory neuron containing magnetite particles. Based on magnetite-containing
structures found in the beak of homing pigeons [10], a model of dipolar interactions
between neighboring superparamagnetic particles was suggested that could account
for how the direction of the geomagnetic field could transmit physiological signals
through interactions with mechanoreceptors in neighboring nerve cells or through
the opening of ion channels in these cells [11]. However, more recent studies suggest
that the magnetite-containing structures in the beak are not sensory neurons, but
macrophages, which are specialized cells involved in the detection and destruction
of bacteria and other harmful organisms [12]. Thus, despite the physical viability
of the mechanism, a structural candidate for a magnetite-based magnetoreceptor is
still missing.

Radical-Pair-Based Magnetoreception
The fact that avian magnetic navigation is light-sensitive led Schulten to propose a
chemical magnetoreception pathway based on the radical-pair mechanism, in which
a photo-induced chemical reaction occurring in a bird’s eye can be influenced by
the direction of the geomagnetic field [13]. Absorption of a photon by a molecule
and its subsequent (photo)reduction by an electron donor molecule nearby produces
a spin-correlated radical pair, with each radical carrying an unpaired electron
with a magnetic spin. Before conversion to a reaction product, the short-lived
radical pair oscillates between singlet and triplet spin states, because of hyperfine
interactions between nuclear spins in at least one of the radicals and the electron
spins. Application of an external magnetic field modulates these oscillations and
influences the reaction product (Fig. 2). The anisotropic shape of the molecules
leads to anisotropic hyperfine interactions, which in turn are the source of magnetic
anisotropy resulting in the reaction product being influenced by the direction of
an externally applied magnetic field. A proof-of-principle laboratory experiment on
an immobilized (frozen) synthetic radical-pair model has shown that radical-pair
recombination time can indeed be influenced by an applied static magnetic field as
weak as the geomagnetic field and that the magnetic field effects are orientation-
sensitive [14]. What is more, the influence of a static field can be modified by the
application of an additional time-varying magnetic field that matches the frequency
of the aforementioned oscillations between singlet and triplet states. Indeed, the fact
that birds were magnetically disoriented when subject to weak radiofrequency fields
speaks for the radical-pair mechanism underlying the inclination compass. The
molecular identity of the radical-pair mechanism in birds has not been elucidated
yet, although members of the cryptochrome protein family (discovered in 1993 by
Ahmad and Cashmore [15] and proposed by Schulten and coworkers [16] to have
a role in the radical-pair mechanism) are currently considered the best molecular
candidates. A comprehensive treatment of the subject is provided in the review by
Hore and Mouitsen [17].
While it is accepted that the radical-pair mechanism is the most probable
explanation for how birds exploit the geomagnetic field during long-distance
1638 N. M. Dempsey

Fig. 2 Illustration of the radical-pair mechanism. The plots are simulations for a simple model
radical pair showing the probability that the radical pair is in its singlet state and how the spin
dynamics are affected by the presence and direction of an Earth-strength magnetic field. The fast
oscillations come from the hyperfine interactions and the slower modulation, most clearly seen in
the red trace, comes from the interactions of the electron spins with the magnetic field. (Courtesy
of P. Hore)

migration (in combination with other cues which exploit their sense of sight, smell,
and hearing), this does not exclude magnetic-particle-based magnetoreception in
birds and other animals. Indeed, the application of a relatively strong magnetic
field pulse, which could reorient the direction of magnetization of single-domain
magnetite particles, was shown to lead to deflected orientation in adult migratory
birds (experienced birds which have an internalized “map” of the regions they
cover during migration). Magnetic-particle-based magnetoreception could explain
the apparent use of local magnetic map information in the late homing phase of
bird migration and also in the behavior of nonmigratory birds and other animals.
For example, some animals, such as the mole rat, build their nests in a direction
determined by the orientation of the geomagnetic field.
Despite significant progress made in studies of how a wide range of animals
are influenced by the geomagnetic field, there is no direct experimental proof
of either the radical-pair mechanism or the magnetic-particle-based mechanism
34 Magnetism and Biology 1639

for magnetoreception. This is also the case for the electromagnetic induction
mechanism proposed to explain magnetoreception in certain fish (sharks, skates,
and rays) known to have highly sensitive electroreceptors. Further advances in the
study of magnetoreception in animals will require a multidisciplinary approach
combining expertise in the fields of (bio)physics and (bio)chemistry, molecular- and
neurobiology, and genetics [18].

Magnetic Fields Produced by the Body

Electrical currents circulating in the bodies of humans and animals create very weak
magnetic fields. These currents are produced in muscle cells during contraction and
in nerve cells during signal transmission. All organs with muscle and nerve cells thus
produce fluctuating magnetic fields, and the largest are associated with synchronous
depolarization of muscle cells in the heart. The largest magnetic fields associated
with nerve cells are produced in the brain, where assemblies of neurons show
synchronous behavior. Typical magnetic field values associated with the human
body are shown in Fig. 3, together with data for the Earth’s magnetic field and
for electromagnetic fields produced by human activity (e.g., power transmission,
wireless communication, etc.), referred to as electromagnetic smog. The maximum
field values produced by various other magnetic field sources, some of which are
used in magnetobiology, are also included in the figure.
The electrical and magnetic signals produced by circulating currents are linked
through the laws of electromagnetism. The well-established techniques to measure
the electrical signals of the heart and brain, electrocardiography (ECG) and elec-
troencephalography (EEG), have their magnetic counterparts, magnetocardiography
(MCG) and magnetoencephalography (MEG). Seeing that the magnetic fields
produced by the body can be orders of magnitude lower than the Earth’s field
and electromagnetic smog, it is not surprising that measuring biomagnetic fields is
very challenging and has lagged far behind electrical measurements. Luigi Galvani
first recorded electrical activity in dissected muscles in 1786, while in 1842 Carlo
Matteucci demonstrated that electrical current pulses accompanied the heartbeat
of a frog. Thirty-five years later, Augustus Waller published the first human
electrocardiogram. Electrical measurement of brain activity was first reported in
1875, when Richard Caton measured electrical currents in the brains of monkeys
and rabbits, and he associated changes in current levels with functional activities
(rotation of the head and mastication) [19]. Fifty-four years later, Hans Berger
published the first EEG which demonstrated electrical potential oscillations in the
human brain at frequencies in the range 8–13 Hz, known as α-waves [20]. But it was
not until 1963 that the first magnetic signals from the human heart were measured
by Baule and McFee, using two coils with ferrite cores and two million windings
each, connected in series [21]. By winding the coils in opposition, they produced a
gradiometer that was able to cancel the background magnetic signals and measure
temporal variations in the magnetic field gradient produced by the heart. Four
years later, David Cohen addressed the background noise problem by carrying out
1640

Resistive coils (cns)


Superconductive coils Pulsed electromagnets
Bitter coils (non-destructive)

10 20 30 40 50 60 70 80 90 100
Human brain Electromagnetic
smog (EMS) Hybrid coils
Muscle Earth’s magnetic field
Heart Pulsed electromagnets
(destructive)

–12 –11 –10 –9 –8 –7 –6 –5 –4 –3 –2 –1 2


10 10 10 10 10 10 10 10 10 10 10 10 1 10 10

Magnetic field intensity (Tesla) NdFeB magnet


Ferrite magnet

Fig. 3 Typical magnetic field values associated with the human body compared to the Earth’s magnetic field and to a range of man-made fields
N. M. Dempsey
34 Magnetism and Biology 1641

Table 1 Complementarity of electrical and magnetic measurement techniques used to study


biomagnetic fields
Electrical Magnetic
Scalar potential Vector field
Electric potential measured relative to a reference Reference-free
electrode
Electrical conductivity varies across matter (scalp, Magnetic permeability practically
skull, cerebrospinal fluid, etc.) constant

Table 2 Brain waves corresponding to neuronal oscillations


Frequency (Hz) <0.2 0.2–3.5 4–8 8–13 14–30 30–90 >90
Brain wave Ultralow Delta Theta Alpha Beta Gamma High-frequency
name oscillations

measurements in an electromagnetically shielded room, and he measured temporal


variations in one component of the field using a single coil [22]. Then in 1968,
he succeeded in measuring the magnetic signal associated with α-waves in the
brain [23]. In 1972, he increased the sensitivity of MEG measurements by several
orders of magnitude by replacing the induction coil with recently developed SQUID
gradiometers. One drawback of commercial SQUID-based MCG and MEG systems
is the need to cool the detectors using liquid helium, so the use of alternative detector
systems based on optically pumped magnetometers [24] and GMR sensors [25] is
being explored. These types of sensors are also being developed for much more
localized measurements, discussed below.
The electrical and magnetic measurement techniques used to study biomagnetic
fields are complementary, and while electrical measurements are the most used,
due to relatively simple and cheap implementation, magnetic measurements offer
some distinct advantages (Table 1). In both cases, arrays of sensors (a few tens
for the heart and up to a few hundred for the brain in research settings) are
used to measure signals from millions of cells, and mathematical tools have been
developed to address the inverse problem of modeling current distributions at the
source that could give rise to the electrical and magnetic fields detected outside the
body. Different brain waves corresponding to neuronal oscillations are classified
according to their frequency, as shown in Table 2. The different types are associated
with different brain functions, and they can act in a cooperative fashion. Brain wave
activity varies depending on the state of consciousness (awake vs different states of
sleep). Specific changes in oscillations measured using EEG and MEG have been
associated with neurocognitive and neuropsychiatric disorders (autism, attention
deficiency hyperactive disorder, schizophrenia), and such measurements are now
being used to study dynamic cognitive neurological processes in research settings
[26]. In clinical settings, MEG is used for preoperative functional mapping and
localization of epileptic foci. MEG signals are typically measured using an array
of up to a few hundred SQUIDs (Fig. 4) within in a helmet-shaped liquid helium
dewar which surrounds the skull.
1642 N. M. Dempsey

Fig. 4 Schematic of an array


of SQUID sensors used in a
commercial
magnetoencephalography
system. (Courtesy of
MEGIN)

EEG and MEG are large-scale sensing techniques, measuring the electrical and
magnetic signals produced by ionic currents in large assemblies of neurons. More
localized measurements using miniaturized sensors or sensor arrays can be used to
measure the signatures of intracellular and extracellular ionic currents from single
neurons or small groups of neurons adjacent to the sensor. Localized measurements
should lead to a better understanding of large-scale measurements and help to
shed light on how the brain works. Various techniques have been developed for
electrophysiological measurements using micro-scaled metallic electrodes or glass
pipettes filled with ionic solution. As with their large-scale counterparts, com-
plementary information can be achieved with micro-scaled magnetophysiological
measurements. The magnetic field sensors should be compact and highly sensitive
and operate at body temperature. A wound toroidal coil has been used to measure the
magnetic field associated with a single giant nerve axon from a dissected crayfish,
threaded through the coil [27]. Nitrogen-vacancy center diamond sensors have been
used to measure the magnetic fields produced by single neurons extracted from
marine worms and squids and from single neurons inside intact worms [28]. GMR
sensors mounted on a needle-like support to create a “magnetrode” have been used
for the in vivo magnetic recording of the neuronal activity of a cat [29].

Iron in the Body

Iron is an essential trace element for most living organisms, ranging from bacteria
to humans. A healthy adult body contains roughly 4 g of iron and requires a daily
dietary intake of 10 or 20 mg for men and women, respectively, only 10% of which is
absorbed by the body. While iron plays a critical role in oxygen transport and redox
processes, its ability to produce free radicals renders it potentially toxic, and a series
34 Magnetism and Biology 1643

Table 3 Distribution of iron in the body [30]


Protein Fraction Localization
Hemoglobin 65% Red blood cells
Myoglobin 4% Muscle tissue
Ferritin 15–30% Throughout the body but concentrated in the liver, spleen, and bone
marrow
Transferrin 0.1% Blood plasma

of complex processes have evolved to maintain body iron levels at a healthy level.
Iron is bound to various proteins within the body (Table 3). The role of hemoglobin
is to transport oxygen in the blood, while that of myoglobin is to carry oxygen to
muscle tissue. Ferritin is a spherical protein cage, which serves to oxidize and store
iron in the form of ferrihydrite nanoparticles (9H2 O.5Fe2 O3 ) of diameter up to 8 nm.
Transferrin serves to transport iron through the blood.
In 1936, Linus Pauling and Charles Coryell measured the magnetic suscepti-
bility of oxygenated and deoxygenated hemoglobin, as well as other hemoglobin
derivatives [31]. They confirmed earlier reports that oxygenated hemoglobin was
diamagnetic and discovered that deoxygenated hemoglobin is paramagnetic and
estimated a magnetic moment of 5.46 μB per iron atom. The paramagnetism is due
to the presence of high-spin Fe2+ ions. While there is still debate about the elec-
tronic structure of iron–oxygen bonds in hemoglobin [32], these bonds play a key
role in oxygen binding and release in red blood cells which transport oxygen from
the lungs to tissue throughout the body. The magnetic susceptibility of red blood
cells is determined by its components, the major ones being diamagnetic water,
diamagnetic oxygenated hemoglobin, and paramagnetic deoxygenated hemoglobin.
The susceptibility difference between completely deoxygenated and completely
oxygenated red blood cells is 2.7 10−7 . When a person is infected by malaria, the
parasite feeds on hemoglobin, converting it to paramagnetic hemozoin crystals [33].
The number and size of hemozoin crystals, which have a distinctive brown color
resulting in hemozoin being known as malaria pigment, depend on the stage of
parasite development. The natural occurrence of iron in the human body is exploited
in a number of applications, as will be detailed below.

Magnetobiology

In this section, we will discuss various ways in which magnetism is exploited in


biology and medicine. Advances made in magnetobiology have benefited from
progress made in the fabrication of magnetic materials and devices. While some
such developments originally targeted other fields of applications such as infor-
mation technology or energy, there is a growing effort in developing materials
and devices specifically targeting biomedical applications. We will begin with a
short description of magnetic nanoparticles used in various biomedical applications.
1644 N. M. Dempsey

Then we will present examples of magnetobiology applications in imaging, fixation,


actuation, and heating. We will discuss effects of magnetic fields on living objects
and will finish up by addressing safety issues.

Magnetic Nanoparticles for Biomedical Applications

Magnetic nanoparticles are used in a range of diagnostic and therapeutic appli-


cations [34]. The vast majority of studies concern iron-oxide particles (γ-Fe2 O3
maghemite and Fe3 O4 magnetite), though other transition metal oxides and metallic
(e.g., Fe, Co) particles are also being explored for use. The relatively high-saturation
magnetization of metallic particles compared to oxide particles leads to stronger
magnetic forces, but metallic particles are prone to oxidation and are toxic. What
are referred to as magnetic particles may have single or multiple nanometer-sized
magnetic cores embedded in a nonmagnetic coating or matrix [35]. The matrix,
which is typically a polymer or organic material, serves to reduce agglomeration
of the constituent magnetic particles and to increase biocompatibility. The matrix
may also serve as an anchor for functional groups, for example, to enable binding
to biomolecular targets or to favor accumulation of particles in a specific organ or
tumor. In the case of multicore particles, the cores may be physically separated by
the matrix or may be in direct contact, in which case they behave as a polycrystalline
single core. Particles in suspension have a surface layer, known as the hydrodynamic
layer in water-based suspensions, which serves as a boundary between the particle
and the liquid it is suspended in. Micron-sized multicore particles are often referred
to as magnetic beads. Magnetic particles are characterized by the size of their
magnetic cores, the relative distribution of cores within multicore particles, the
diameter of the matrix, and the thickness of the functional and boundary layers
[35]. When the magnetic cores are sufficiently small, they show superparamagnetic
behavior.
Iron-oxide nanoparticles with certain sugar-based or silica-based coatings are the
only type of magnetic nanoparticles approved for clinical use by agencies such as
the European Medicines Agency (EMA) or the US Food and Drug Administration
(FDA) (Table 4) [35]. While our principle concern here is the exploitation of the
particles’ magnetic properties, one of the prescribed uses of iron-oxide nanoparticles
is to boost iron levels in anemic patients. Indeed, it is the fact that the body
naturally contains iron and is capable of biodegrading iron-oxide nanoparticles
through iron metabolism that renders these particles biocompatible. Depending on
the application, iron-oxide nanoparticles are administered to the body by injection
into the bloodstream or into tissue (e.g., a tumor), by inhalation, or by swallowing.
As soon as these foreign objects enter the body, the immune system sets about
recognizing and eliminating them. What happens to the nanoparticles depends
on their size, shape, coatings, charge, and mode of administration. The amount
and distribution pattern of the nanoparticles throughout the body is referred to
as their biodistribution, while the rates of their recognition and removal by the
immune system, metabolism, and excretion from the body is referred to as their
34 Magnetism and Biology 1645

Table 4 Iron-oxide magnetic nanoparticles approved for administration to humans [35]


Usage Trademark Coating
MRI contrast agent Resovist® Carboxydextran
Iron replacement agent Feraheme® Carbohydrate
Sentinel node detection agent Sienna+® Dextran
Magnetic thermoablation agent NanoTherm™ Aminosilanes

Fig. 5 Iron-oxide nanoparticle (IONP) biodegradation and general iron transport and metabolism
pathway in the body (MPS mononuclear phagocytic system) [30]. (With kind permission from
Royal Society of Chemistry)

pharmacokinetics [30]. Very fine particles (<10–15 nm) get filtered through the
kidney, while larger particles (>100 nm) are typically trapped in the liver and spleen.
The time taken for the concentration of nanoparticles injected into the blood to
decrease to half its initial value is known as the blood half-life. It can be lengthened
by modifying the surface of the nanoparticles so as to hide them from the immune
system, by attaching them to red blood cells, or by tricking the immune system by
injecting decoy particles. Blood half-life times vary from minutes to several days in
rodents and from 1 h to 1 day in humans.
Once trapped, exposure of the nanoparticles to acidic environments and biochem-
ical reactions leads to their biodegradation (Fig. 5). Degradation rates depend on the
particles’ location, their coatings, and their initial dosage. Iron from nanoparticles
can be stored in the iron protein complexes ferritin and transferrin, and transferrin
1646 N. M. Dempsey

can be transported through the blood to bone marrow to make hemoglobin or to


muscles to make myoglobin. Some nanoparticles and by-products of biodegradation
simply get excreted from the body. Biodegradation leads to a modification of the
magnetic properties of the injected iron-oxide nanoparticles, and magnetization
measurements together with high-resolution imaging were used to study the process
in vivo [36].
Toxicity studies of iron-oxide nanoparticles and their degradation products in
animal models have revealed potential impact on blood chemistry, gene expression,
and kidney, liver, and heart function. In human toxicity tests, minor side effects
including back pain, itchiness, and headaches were reported by a quarter of patients,
while severe adverse effects (e.g., chest pain, shortness of breath, skin rash, hypoten-
sion) were reported by less than 0.5% of patients. Considering the complexity
of the human body and the wide range of parameters characterizing magnetic
nanoparticles, further studies are obviously needed to qualify the dose-dependent
biocompatibility of specific magnetic nanoparticles. Finally, the modification in
magnetic properties caused by biodegradation affects their functional properties,
and thus the rate of degradation must be studied and if possible tailored to suit the
specific application being targeted [37].

Magnetic Imaging

Magnetic imaging techniques are complementary to other medical imaging tech-


niques based on the use of potentially hazardous X-rays (X-ray imaging, computed
tomography [CT]) or ionizing radiation (positron emission tomography [PET]).
Here we will describe three magnetic imaging techniques, namely, magnetic
resonance imaging (MRI), functional magnetic resonance imaging (f-MRI), and
magnetic particle imaging (MPI). While MRI is a very well-established technique
used in hospitals and clinics around the world, f-MRI is mostly used in neurocogni-
tive research laboratories while MPI is still in the development stage.

Magnetic Resonance Imaging (MRI)


Magnetic resonance imaging exploits spatially resolved nuclear magnetic resonance
(NMR) of specific nuclei in the body. By far, the greatest number of MRI studies
concerns hydrogen nuclei and serves to map the location of water and fat. The
detected signal depends on the environment of the nuclei, and thus image contrast
allows different types of tissue to be distinguished. The first MRI images were
published by Paul Lauterbur in the early 1970s, who imaged thin tubes of water
[38] and then a living mouse [39]. Advances in the imaging technique made soon
after by Paul Mansfield led to much faster acquisition times and clearer images [40].
Lauterbur and Mansfield were jointly awarded the 2003 Nobel Prize in Physiology
or Medicine for their “discoveries concerning magnetic resonance imaging.”
In MRI, a strong static homogeneous magnetic field is used to align the nuclear
spin moments of hydrogen or other nuclei. A radiofrequency (RF) magnetic field
34 Magnetism and Biology 1647

is then applied to induce resonance between the Zeeman-split nuclear energy


levels, which perturbs the slight alignment of the nuclear spins with the static
field. When the excitation field is turned off, the nuclear spins precess back
toward the applied field, producing an RF signal, which can be detected using
a pickup coil. Gradient coils are used to produce linear variations in the main
magnetic field in three directions, causing the resonance frequency of protons to
vary as a function of position. Spatial localization of the RF signal to build up
2D and 3D images can then be achieved based on differences in the frequency,
phase, and timing of the signal, as well as its distance from the receiver coil.
Relaxation occurs by different processes and is characterized by two different time
constants. Longitudinal relaxation caused by spin–lattice interactions is quantified
by T1 , which characterizes the time taken for the nuclear spins to regain their
thermal equilibrium alignment with the static field. Transverse relaxation caused
by spin–spin interactions is quantified by T2 , which characterizes the time for phase
decoherence between the perpendicular components of the nuclear spins. The T1 and
T2 values depend on tissue type. Different types of images are produced by varying
the time between applied RF pulses and the time between the delivery of an RF
pulse and the recording of the RF response signal. Examples include T1 -weighted
and T2 -weighted images, in which the contrast and brightness are predominately
determined by T1 and T2 properties of tissue, respectively. Contrast agents can
be introduced into the body to enhance the quality of MRI images. Paramag-
netic contrast agents (e.g., complexes containing Gd3+ ) interact with neighboring
hydrogen nuclei and primarily affect T1 , while superparamagnetic (e.g., iron-oxide
nanoparticles) contrast agents strongly perturb the local magnetic field and primarily
affect T2 .
MRI is used for general anatomical imaging, cardiovascular imaging (detailed
assessment of heart and blood vessels), and musculoskeletal imaging (cartilage,
joints, soft tissue, spine). Real-time MRI is used for continuous imaging of moving
objects (e.g., the heart), guidance during targeted drug delivery [41], and in vivo
tracking of immune cells [42]. Interventional MRI is used to guide minimally
invasive procedures, while intraoperative MRI is used during surgery.

Functional Magnetic Resonance Imaging (f-MRI)


When a region of the brain is activated, local oxygen consumption increases,
resulting in a buildup of deoxygenated hemoglobin. Increased blood flow in the
region is then triggered, resulting in increased oxygenated hemoglobin levels.
While oxygenated hemoglobin is diamagnetic and magnetically indistinguishable
from brain tissue, deoxygenated hemoglobin is paramagnetic and locally modifies
magnetic field gradients. Functional magnetic resonance imaging (f-MRI) measures
brain activity by following either changes in blood oxygen content using blood-
oxygen-level-dependent (BOLD) contrast [43] or changes in cerebral blood flow
(CBF), with the former being by far the most widely used approach [44]. f-MRI
is now extensively used in neuroscience research laboratories and some clinical
settings, where brain activity is monitored in the rest state, or in an active state
1648 N. M. Dempsey

following motor, sensory, or emotional stimulation. It is often used in combination


with EEG or MEG, which directly measure the electrical and magnetic signals
associated with brain activity. f-MRI is used in the study of psychiatric and neuro-
logical diseases (schizophrenia, depression, autism, drug addiction, etc.), to monitor
certain therapies and to study pharmacologic efficacy [44] as well as functional
gastrointestinal disorders (irritable bowel syndrome and functional dyspepsia) [45].
However, cognitive studies still face many challenges in interpretation. f-MRI is also
used in preoperative planning and intraoperative guidance in brain surgery (epilepsy
treatment or tumor removal) [46].

Static Magnetic Field Sources Used for MRI and f-MRI


The signal strength and spatial resolution achievable in MRI and f-MRI depends
directly on the intensity and homogeneity of the static field applied to align the
nuclear spins. Most whole-body MRI systems now use closed-bore superconducting
coils (Fig. 6), with hospital- and clinic-based systems typically operating in the
range of 1.5–3 T [47]. Open-bore MRI systems have also been developed for use on
patients that cannot be imaged in a closed-bore system (e.g., claustrophobic or obese
patients) and to allow interventional MRI [47]. Characteristics of clinically used

Fig. 6 Image of an MRI suite with a whole-body MRI system based on a closed-bore supercon-
ducting coil together with an MRI image of a human head. (© Halkin Mason Photography)
34 Magnetism and Biology 1649

Table 5 Characteristics of static magnetic field sources used for MRI


MRI system Field geometry Field source Field strength
Whole-body Solenoid Superconducting coil 1.5–3 T
Closed-bore
Open-bore Dipolar Superconducting or resistive coil 0.5–1.2 T
Open-bore C-shaped NdFeB permanent magnets 0.2–1.0 T

MRI static field sources are compared in Table 5. In all cases, shimming (passive or
active) is used to increase the field homogeneity, and shielding is used to limit the
stray field produced by the magnet.
Whole-body systems with 7–9.4 T fields are being used for f-MRI in neuro-
science studies in a few research laboratories and university hospitals, while systems
with even higher field values are being developed in very specialized laboratories
[48, 49]. The increase in field strength and field gradient applied during imaging
leads to increased sensitivity and spatial, temporal, and spectral resolution and
allows imaging of a range of nuclei (e.g., 13 C, 17 O, 31 P, 23 Na, 35 CI, and 39 K)
to study energy metabolism and electrolyte physiology in the brain. However, the
very high installation and running costs associated with mid- and high-field whole-
body MRI systems limit their use to well-funded hospitals, clinics, and institutes.
Furthermore, the large size of such systems and the need for cryogenic liquids
to cool the superconducting coils restricts their use to very specific environments.
Small-bore systems based on permanent magnets and resistive or superconducting
coils are being used for imaging human limbs in doctors’ surgeries [50], as well as
primates and small animal models (mice, rats) in research laboratories. MRI systems
with fields of just a few tens of mT are being developed with the aim to expand
the use of MRI to in-field use (doctors’ surgeries, sports grounds, battlefields, etc.)
and to resource-poor environments. Approaches to compensate for the very low
signal-to-noise ratio achieved with such low fields include the application of a strong
inhomogeneous magnetic field pulse to polarize the sample followed by readout in a
low field (pre-polarized MRI) and the use of very fast imaging techniques combined
with sparse sampling strategies [51].

Magnetic Particle Imaging (MPI)


Magnetic particle imaging (MPI) is an emerging medical imaging technique, first
reported in 2005, that exploits the nonlinearity of the magnetization curves of
superparamagnetic magnetic nanoparticles and the fact that their magnetization
can be saturated in a given field [52]. While magnetic nanoparticles can be used
as contrast agents to enhance MRI images, they are the only source of signal in
MPI. MPI maps the spatial and temporal distribution of a magnetic tracer (the
magnetic nanoparticles), in the same way that PET maps the temporal distribution
of a radioactive tracer. Compared to radionuclides, magnetic nanoparticles are safer
and cheaper and have a longer shelf life.
1650 N. M. Dempsey

Both time-varying and static magnetic fields are used during MPI (Fig. 7) [52].
An oscillating magnetic field is used to flip the direction of magnetization of the
magnetic nanoparticles, and pickup coils are used to record their M(t) response. As
the magnetization curves are nonlinear, a Fourier transform of the induced signal is
characterized by the excitation frequency and its higher harmonics. The presence of
these harmonics allows us to distinguish between the tracer signal and the excitation
field signal. To restrict the region in which magnetic nanoparticles are excited, a
static magnetic field of sufficient intensity, with either a field free point (FFP) or
a field free line (FFL), is superimposed on the oscillating magnetic field [53]. The
static field serves to magnetically saturate all magnetic nanoparticles outside the
FFP or FFL, so that they do not respond to the oscillating magnetic field. The
position of the FFP or the FFL is scanned over the total area to be imaged, to
build up a 3D picture of the concentration distribution of magnetic nanoparticles.
Scanning of the FFP or FFL can be achieved either mechanically or with the use
of additional magnetic field coils. The typical magnetic field frequencies applied
in MPI systems are in the range of 10–100 kHz, while their peak amplitudes are
in the range 10–100 mT. The sensitivity of the technique depends on the magnetic
moment of the nanoparticles, while the spatial resolution depends on their magnetic
susceptibility as well as the gradient of the static field. MPI offers exceptional
spatial (mm) and temporal (ms) resolution. While magnetic nanoparticles developed
as MRI contrast agents have been used as MPI tracer agents, the optimum
properties are not the same for both techniques, and much effort is going into
developing magnetic nanoparticles specifically for MPI. Progress is also being
made in image reconstruction and scanner design, with the aim to further improve
sensitivity and time resolution and to upscale to full-body systems for clinical
use. MPI is used to image the flow of blood containing magnetic nanoparticles in
cardiovascular studies, to track specific cells labeled with magnetic nanoparticles,
and to monitor the position of interventional instruments coated with magnetic
nanoparticles [53, 54].

Magnetic Fixation

A simple but important use of mm-cm-sized magnets in the field biomedical


applications concerns fixation. The high energy densities of Nd2 Fe14 B- and SmCo5 -
based magnets give rise to high retention forces with relatively small magnets. A
number of applications exist in dentistry and related fields [55]. Magnets are used
to hold removable overdentures in place. In orthodontics, their role goes beyond
fixation, as they are used to apply forces to induce tooth movement. Magnets are
also used for the retention of maxillofacial prostheses, designed for the rehabilitation
of patients with defects or disabilities that were present when born or developed
due to disease or trauma. Magnets are also used for temporary anchoring of tools
such as cameras, retractors, and robotic cauterizers during minimally invasive intra-
abdominal surgery, so far carried out on animal models [56].
34 Magnetism and Biology 1651

Fig. 7 Schematic of the basic principles of MPI: (a) An oscillating magnetic field (modulation
field, H, green curve) is applied to the magnetic material at a given frequency f1 . As the
magnetization curve (M, black curve) is nonlinear, the resulting time-dependent magnetization (red
curve) exhibits higher harmonics, as shown in the Fourier transformed signal (S, red bars). (b) A
time-independent field is added to the modulation field. The oscillating field does not significantly
change the magnetization of the material, as it is always in saturation. In this state, harmonics of
the oscillating field are almost nonexistent. The gray box indicates those harmonics used for image
formation [52]. (With kind permission from Springer Nature)
1652 N. M. Dempsey

Magnetic Actuation

Magnetic fields can be used to move objects. A homogeneous field B can be used to
apply a torque (τ ) to an object of magnetic moment (m), oriented at an angle to the
field, causing it to rotate. The torque is given by

τ = m × B (1)

Units of m are Am2 , and units of τ are Nm. An inhomogeneous magnetic field
can pull or push the object via the field gradient force,

f = ∇ (m . B) (2)

where ∇B is the magnetic field gradient. Usually, the object does not have a
permanent magnetic moment; the moment is induced by the applied field. When
the response is linear, m = χvB/μ0 , where χ is the dimensionless susceptibility of
the object and v is its volume. Units of f are N. Since the magnetization M = m/ v,
the force density F in Nm−3 can be written as

F = (1/2 μ0 ) χ ∇B 2 (3)

The magnetic field gradient force attracts ferromagnetic, ferrimagnetic, super-


paramagnetic, and paramagnetic objects, which are characterized by a positive
susceptibility, toward high-field regions while it repels diamagnetic (and super-
conducting) objects, which have negative magnetic susceptibility, away from these
regions.
Magnetic actuation involving rotation or displacement is used in a wide range
of bio- and biomedical applications, some of which are listed in Table 6. The vast
majority of biological targets (cells, proteins, DNA, etc.) are weakly diamagnetic.
To easily manipulate such diamagnetic targets with a magnetic field gradient,
they are typically labeled with magnetic micro- or nanoparticles, which pull them
along. Particles showing superparamagnetic behavior are normally used to avoid
agglomeration in the absence of the external magnetic field and to ensure that the
targets can be released on demand.
The magnetic particles may be internalized by phagocytosis (when cells eat
particles from the surrounding medium), electroporation (when an electric field
is applied to cells in order to render the cell membrane permeable – a magnetic
version of this will be discussed below), or direct injection, in the case of sufficiently
large targets, such as embryos. Alternatively, they may be attached to the surface
of the biological target using electrostatic interactions, ligand–receptor complexes
(e.g., biotin–streptavidin), and adhesive molecular coatings (e.g., poly-ornithine).
When very high fields or field gradients are used, repulsion between a diamagnetic
biological target and a magnetic field source can be exploited to manipulate
34 Magnetism and Biology 1653

Table 6 Examples of the use of magnetically induced actuation in fundamental studies in biology
and biomedical applications
Actuation Application
Magnetic trapping (non-lab-on-a-chip) Diagnostics and therapeutics
Magnetic trapping, transportation, mixing of Chemical and biological analysis diagnostics
cells in lab-on-a-chip devices
Magnetic tweezers Measurement of physical properties,
mechanical stimulation
Magnetically induced cell destruction Cancer treatment
Magnetoporation Payload delivery to cells
Magnetic orientation Tissue engineering

the target. Levitation can be achieved when the magnetic force compensates
gravity:

F = ρg (4)

where ρ is the density of the target. A Bitter coil producing a magnetic field of 16
Tesla was used to levitate a living frog [57], while Nd-Fe-B micromagnets producing
a maximum field gradient of 106 T/m were used to levitate living cells [58].
Deoxygenated red blood cells are paramagnetic and have been manipulated directly,
without being magnetically labeled. Finally, some biological targets naturally con-
tain ferrimagnetic iron-oxide nanoparticles (e.g., magnetotactic bacteria, splenic red
pulp macrophages, etc.), which renders them easy to manipulate by a magnetic field.
As the magnetic field source used to achieve actuation, one can use current-
carrying wires and coils or permanent magnets (Fig. 3). The field strengths
achievable with resistive wires and coils operated in continuous mode are typically
restricted to values of the order of tens to hundreds of mT because of Joule heating,
while fields of the order of 1–10 T can be achieved in pulsed mode. Superconducting
coils can produce fields of this strength in continuous mode, but they require cooling
and are both cumbersome and expensive, and thus they are impractical for benchtop
magnetic actuation in biology laboratories. Permanent magnets based on hard ferrite
(e.g., Ba2 Fe12 O19 , Sr2 Fe12 O19 ) and rare earth transition metal (e.g., Nd2 Fe14 B,
SmCo5 ) phases produce stray fields of the order of a few hundred mT and 1 T,
respectively.
Miniaturization of the magnetic field source may be desirable for two reasons.
Firstly, it allows to locally address and manipulate small-scale biotargets. Secondly,
reducing the size of the field source while maintaining its intensity leads to an
increase in the magnetic field gradient it produces and thus increases the force it can
apply on a biotarget. When scaling down the size of a coil, it is usually necessary
to reduce the current and thus the field intensity, because of increased heating.
Nevertheless, planar coils suffer this less than wire coils, as their high surface-to-
volume ratio allows for improved cooling [59]. Micro-scaled permanent magnet
1654 N. M. Dempsey

field sources can be fabricated using hard magnetic powders [60, 61] or by micro-
patterning hard magnetic films using topographic [62] or thermomagnetic [63]
techniques. When miniaturized field sources are required, it is also possible to use
micro-scaled soft magnets based on Fe, Ni, or Co and their alloys and to magnetize
them using larger coils or permanent magnets. The choice of field source depends
on the specific application. Micro-coils and soft micromagnets are very suitable
when easy control of the force intensity is required, while hard micromagnets are
particularly well adapted when space is limited.

Magnetic Trapping (Non-Lab-on-a-Chip)


High-gradient magnetic separation is used for garbage sorting, mineral processing,
bioprocessing (e.g., extraction of proteins from fermentation broth, plasma, milk,
whey extracts) [64], and water purification [65]. It is extensively used for the
separation of biological entities in diagnostics, and its use in various therapeutic
applications is being studied, as detailed hereafter.

Magnetic Cell Separation


In vitro trapping of magnetically labeled cells can be achieved by simply approach-
ing a bulk permanent magnet to the sample holder. Such trapping is used during
washing or changing of the cell medium or separation of labeled cells from non-
labeled cells in single sample holders or multi-well plates. The efficiency of cell
trapping can be increased by passing the solution through a column containing fine
magnetic wire wool or magnetic beads, with external magnets placed next to the
column [66]. The magnets serve to magnetize the wool or beads, resulting in high
magnetic field gradients close to the cells. Automated high-throughput magnetically
activated cell sorting (MACS) systems can be commercially sourced, together with
magnetic nano- or microparticles biochemically functionalized to attach themselves
to specific cell types.

Magnetic Hemofiltration
High-gradient magnetic separation has been proposed for hemofiltration, in which
an extracorporeal device is used to remove targeted entities from whole blood in
a reservoir [67] or directly from blood drawn from and repumped into the body
[68] (Fig. 8). This approach could be used to eliminate undesirable entities such as
pathogens responsible for sepsis, toxins due to kidney or renal failure, circulating
tumor cells, viruses, and radionuclides or to harvest desirable entities such as stem
cells, for therapeutic use. A magnetic trap has been added to a dialysis circuit to
perform magnetically assisted hemodialysis [69]. While much work is going into
developing biofunctionalized magnetic particles suitable for binding to a wide range
of specific blood-borne targets [68], malaria-infected paramagnetic red blood cells
do not require labeling with magnetic particles to be magnetically extracted from
blood [70].

Magnetic Targeted Drug Delivery


In vivo magnetic trapping of drugs functionalized with magnetic particles is being
studied for targeted drug delivery [34, 71]. The motivation here is to reduce the risk
34 Magnetism and Biology 1655

Fig. 8 Schematic of a magnetic hemofiltration system exploiting high-gradient magnetic separa-


tion (HGMS) to remove targeted entities from whole blood. (Adapted from [68])

of the chemotherapy agent killing healthy cells, by trapping them at a specific site on
a tumor. Efficient trapping allows the overall drug dose administered to the patient
to be reduced. A range of biocompatible magnetic carriers have been developed
using both nano- and micro-sized magnetic particles coated by or embedded in
polymer or silica. The carriers are typically administered by injection in the form
of a ferrofluid into the bloodstream. Magnetic targeting is achieved using external
electromagnets, external or implanted bulk permanent magnets, implanted soft
magnetic microwires or microrods magnetized by an external field [71], or so-called
magnetic resonance navigation (MRN). In the latter technique, the static field of a
clinical MRI scanner is used to magnetize the magnetic carriers, and the software-
controllable magnetic gradient fields of the system are used to navigate the carriers
in real time along a pre-planned trajectory from their injection site to the targeted
area [72]. Payload release may be activated by changes in pH or temperature in
the targeted area or through thermal, electrical, or magnetic stimulation [73]. The
general approach has also been proposed for targeted delivery of radionuclides and
therapeutic genes [74]. Magnetic drug delivery trials have been carried out on many
animal models, but very few trials have been carried out on humans. Challenges
faced for clinical implementation include the risk of blocking blood vessels through
the accumulation of magnetic carriers, potential toxicity of the magnetic carriers
themselves, and the limited spatial extension of the magnetic field produced by non-
MRN external field sources, which restricts potential applications of such sources
to shallow targets.

Magnetic Manipulation in Lab-on-a-Chip


The lab-on-a-chip (LOC) or micro-total analysis system (μ-TAS) approach to
diagnostics, which is based on the use of microfluidic channels for liquid handling,
is attractive because only very small quantities of samples and reagents are required,
reaction times can be relatively fast, and the large surface-to-volume geometry
is advantageous for surface-based assays. In contrast to large automats used for
1656 N. M. Dempsey

high-throughput screening in diagnostic laboratories, LOC devices are well suited


to point-of-care applications. While electric field manipulation has long been used
in LOC, magnetic manipulation for the trapping, transportation, and mixing of cells
or other biotargets (proteins, DNA, etc.) has emerged [75, 76]. Studies carried out on
magnetic beads, rather than magnetic beads or particles internalized by or attached
to cells, serve as intermediate studies, which should contribute to the magnetic
manipulation of biotargets.

Magnetic Trapping and Separation


Magnetic trapping in microfluidic devices has been used to isolate magnetic beads
from nonmagnetic beads and various types of magnetically labeled cells from non-
labeled cells using external bulk permanent magnets [77] or integrated field sources
based on micro-coils [78], soft micromagnets [79], or permanent micromagnets
[80, 81] (Fig. 9A). Continuous-flow separation of magnetic beads of different sizes
and magnetic susceptibilities has been achieved using a laminar flow chamber with
multiple inlet and outlet channels and an asymmetrically positioned external bulk
magnet (Fig. 9B) [82]. Continuous-flow separation of labeled cells from whole
blood was achieved by flowing the mixture through a channel with magnetic stripes,
misoriented with respect to the flow direction [83]. Magnetic trapping is used
in both immunoassays, which detect the presence of specific molecules such as
proteins through the use of antibodies or antigens, and nucleic acid assays, which
detect genetic material. Magnetic beads act as substrates or solid supports for the
biochemical reactions involved in these bioassays. Magnetic trapping is typically
combined with optical, electrochemical, or magnetoresistive detection techniques.
In the latter case, the magnetic bead serves both as a force vector for movement and
a label for detection [84], while in the case of optical detection, the magnetic bead
may also serve a dual purpose if it is fluorescently marked.

Fig. 9 Schematic image of a microfluidic chip with an integrated array of permanent micromag-
nets used to separate magnetically labeled cells from non-labeled cells [81] With kind permission
from Elsevier (A). Schematic image of a free-flow magnetophoresis chamber used to separate
magnetic particles from nonmagnetic material, according to their size and magnetic susceptibility.
(Reprinted with permission from [82]. Copyright (2004) American Chemical Society (B))
34 Magnetism and Biology 1657

Magnetic Transportation
In most microfluidic systems, fluid flow is used to transport objects within the
channel, but when the objects to be moved are magnetic, then magnetic forces can be
used to help control their transportation [76]. This was first demonstrated by simply
moving an external permanent magnet below a channel containing magnetic beads.
Arrays of sequentially excited electromagnets positioned along a channel can also
be used to drag magnetic beads along, and more complex arrays of independently
addressed current-carrying wires can be used for magnetically controlled transport.
Application of an external magnetic field to further magnetize the beads can
significantly increase the magnetic forces. The risk of heating potentially limits
the use of current induced magnetic fields for magnetic transportation. Arrays of
micromagnets can produce a periodic potential energy landscape in which magnetic
beads are trapped. Superposition of a rotating external magnetic field on such an
array can then be used to translate the energy landscape and thus transport magnetic
beads in a process known as travelling wave magnetophoresis. This process can
be used to separate beads of different sizes. Advances made in the sequential and
parallel, timed magnetically induced movement of an ensemble of single particles
and cells hold great potential for a wide range of applications in cellular biology and
biotechnology [85].

Magnetic Mixing
Magnetic mixing in microfluidic channels of dyes and buffer solutions has been
achieved using bar-shaped stirrers made of permalloy or chains of magnetic particles
rotated under the influence of an external rotating magnetic field [75]. Pumping and
mixing of ionic fluids has been demonstrated using magnetohydrodynamic micro-
pumps in which an electric field and a magnetic field are applied perpendicular to
each other across the height and width of a microchannel. Coupling between the
electric force produced on ions by the electric field and the Lorentz force produced
by the magnetic field results in movement of the liquid along the channel.

Magnetic Tweezers
Magnetic tweezers are tools used by biologists to apply a mechanical force on a
target object. The basic idea is to apply an external magnetic field to a magnetic
particle within or attached to the biological target. Originally developed to measure
the mechanical properties of cells and molecules, the approach is now also used to
study the biological response to mechanical stress.

Measurement of Physical Properties


In 1949, Francis Crick studied the rheological properties of the cytoplasm of cells
(the material within the cell, excluding the nucleus) by observing the motion of
magnetic particles within the cell, under the influence of a magnetic field [86]. He
used colloidal magnetite and elongated oxidized iron particles, their diameter being
one third of their length (2–3 μm), which were phagocytosed by the cells. Particles
were twisted by rotating a homogeneous applied field or dragged through the
cytoplasm with an inhomogeneous field. This relatively crude study demonstrated
1658 N. M. Dempsey

Fig. 10 Schematic of the


magnetic tweezers technique
used to characterize the
mechanical properties of
DNA molecules. (From
Wikipedia, courtesy of Daniel
Förster)

that the cytoplasm has viscous properties and the order of magnitude of the modulus
of rigidity could be estimated. In a twist of fate, the approach was later adapted to
directly measure the elastic properties of DNA molecules! One end of a molecule
was chemically attached to a glass slide, the other to a superparamagnetic bead, and
the ensemble was placed in a microfluidic chamber with external movable magnets
(Fig. 10). By recording the movement of the tethered beads as a function of the
magnetic and applied hydrodynamic forces in the range 10−14 –10−11 N, it was
possible to establish force versus extension, stress–strain curves characteristic of
the DNA molecule. The approach was further refined to measure the elasticity of
a single supercoiled DNA molecule, by using permanent magnets [87] and later
on an array of electromagnets [88] to precisely translate or rotate a bead attached
to the end of a molecule anchored to a glass slide. The fact that a torque can be
applied to a superparamagnetic bead indicates some magnetic anisotropy, which is
attributed to a nonuniform distribution of superparamagnetic nanoparticles within
the bead [89]. Magnetic tweezers have two specific advantages compared to optical
tweezers: the first is that the angular position of the target can be controlled,
and the second is that the biomaterial being manipulated is not exposed to
photodamage.
Magnetic tweezers have also been used to measure the rate of dissociation
of ligand–receptor bonds [90], notably the prototypical ligand–receptor pair,
streptavidin–biotin, commonly used to attach magnetic nano- and microparticles to
biotargets. Magnetic beads were functionalized with streptavidin, and biotinylated
proteins were adsorbed to the surface of a plastic slide. The beads were allowed to
contact the surface so that ligand–receptor binding occurred and then a permanent
magnet was used to pull the receptors away from the ligands. Massively parallel
(100–1000) measurements were made on large batches of single complexes
perturbed by constant forces. Technical advances made in the magnetic beads,
field sources, and force characterization protocols now allow the characterization
of the mechanical properties of a large number of biological samples, from
34 Magnetism and Biology 1659

individual molecules to intermolecular bonds to whole cells using magnetic


tweezers [91].

Mechanical Stimulation
Living cells are able to feel, respond, and adapt to the mechanical properties of their
environment; the process by which cells convert mechanical signals into biochemi-
cal signals is called mechanotransduction. Cells are sensitive to both the spatial and
temporal signatures of mechanical stimuli, and thus to study mechanotransduction,
it is essential to stimulate cells with mechanical cues controlled both spatially and
temporally. Mechanical stimuli have been applied to cells using both direct methods
such as atomic force microscopy and indirect methods including optical or magnetic
[92] tweezer manipulation of beads attached to the cell. The magnetic tweezer-based
approach has the advantage of being minimally invasive and relatively simple to
implement. Nevertheless, attachment of a magnetic bead to the surface of a cell
can potentially perturb the cell due to direct mechanical interactions between it
and the stiff bead. Substrates made of micropillars that can be actuated with a
magnetic field were used to apply local and dynamic mechanical stimuli on cells
adhering to the surface (Fig. 11a) [93]. When an external magnetic field is applied,
pillars containing magnetic nanowires apply a force on cells, while nonmagnetic
posts deflect in response to, and therefore measure, traction forces of the cells.
To overcome the fact that the topology of such discrete surfaces can affect the
cellular behavior, flat soft elastomer substrates containing iron micropillars have
been developed (Fig. 11b) [94]. These can be locally and dynamically deformed
in a magnetic field, and localized deformation can be quantified by measuring the
displacement of fluorescent markers encrusted under the surface of the elastomer
using traction force microscopy. Cells spread on such magneto-active substrates
can be mechanically stimulated both in tension and in compression.
Magnetic actuation of magnetic particles on the surface of cells can be used
to activate mechanically sensitive ion channels, which control the passage of ions
through the cell membrane [95]. Specific binding of magnetic nanoparticles can be
used to activate specific ion channels. Such nanomagnetic actuation can be used to
study biochemical pathways associated with mechanical activation and to evaluate
ion-channel kinetics, and it has potential applications in tissue engineering and
regenerative medicine [95].
Magnetic forces can also be used to study mechanotransduction at the embryo
level. The role of mechanical forces during a specific stage in embryo development
known as gastrulation, during which the first coordinated movement of cells occurs,
has been studied in zebra fish and fruit flies [96, 97]. Magnetic nanoparticles
encapsulated in liposomes, a type of carrier vessel, were injected into specific
regions of embryos that had been genetically modified to arrest development at
the pre-gastrulation stage. Static magnetic forces produced by a ring of hard
magnetic particles and dynamic magnetic forces produced by an array of Fe
micromagnets magnetized by an external electromagnet were used to mechanically
initiate gastrulation in the zebra fish and fruit flies, respectively [96, 97]. These
results demonstrate the mechanosensitivity of biochemical pathways active during
gastrulation.
1660 N. M. Dempsey

Fig. 11 Schematic diagrams of microfabricated magnetically active polymer substrates used for
mechanotransduction studies at the cellular level [93]. (With kind permission “Copyright (2007)
National Academy of Sciences, U.S.A.” [94], Creative Commons CC BY license)

Magnetically Induced Cell Destruction


Magnetically induced mechanical forces produced by low-frequency magnetic
fields on magnetic particles can be used to kill cells. The first report of this
effect concerned the use of 4.5 μm diameter beads containing superparamagnetic
nanoparticles, biofuntionalized to attach themselves to cells [98]. When stimulated
ten times at 5 second intervals with unipolar field pulses of intensity 2 T and duration
150 μs, cell viability was significantly reduced, as a result of cell rupture. The
use of lithographically defined permalloy disks of diameter 1 μm with spin-vortex
ground states, biofuntionalized to attach themselves to cancer cells, was then studied
[99]. In this case, in vitro application of AC magnetic fields of just a few mT in
intensity, at frequencies of a few Hz, was shown to damage cell membranes and
to activate apoptosis (programmed cell death) as schematized in Fig. 12. The use
of low-frequency magnetic fields ensures negligible heating. A number of in vitro
studies using low-frequency fields of relatively low intensity have since been carried
out on different types of cancer cells using synthetic antiferromagnetic microdisks
[100], spherical- and spindle-shaped superparamagnetic particles [101], and iron-
doped carbon nanotubes [102]. Rotating fields of the order of 1 T have been used
to release synthetic antiferromagnetic microdisks internalized by neural stem cells
and to subsequently kill brain cancer cells that internalize the released microdisks
[103]. The neural stem cells used can cross the blood–brain barrier and could thus
34 Magnetism and Biology 1661

2 R

1
Surface

IL
Membrane

ti-
receptors

n
-a
Integrity loss
D
M

10 min
Phospholipid
Nucleus
membrane DNA damage

a.c. magnetic field Tens of hertz

Fig. 12 Schematic diagram of targeted magneto-mechanical cancer-cell destruction using disk-


shaped magnetic particles attached to the cell membrane [99]. (With kind permission from Springer
Nature)

potentially serve as shuttles to safely transport magnetic microdisks to targets within


the brain. In vivo destruction of cancer cells containing vortex microdisks injected
into the brain of mice has been achieved by placing the head of the mice into a
rotating Halbach array of Nd-Fe-B magnets [104].

Magnetoporation
Just as magnetic fields can be used for targeted drug delivery at the macroscale, they
can also be used for delivery at the cell scale in a process known as magnetoporation.
A specific form of magnetoporation which is used to deliver nucleic acids attached
to magnetic nanoparticle vectors into cells is known as transfection. Both static and
time-varying magnetic fields have been used for in vitro and in vivo transfection
of various cell types [105, 106]. The application of static inhomogeneous fields,
typically produced by bulk magnets with a maximum applied field of about 0.5 T,
is believed to increase transfection by rapidly accumulating nucleic acid–magnetic
nanoparticle complexes on the surface of target cells. The higher efficiency of time-
varying magnetic fields, produced by pulsed electromagnets delivering a maximum
applied field in the range of 30–300 mT or vibrating arrays of permanent magnets,
is attributed to the influence of magneto-mechanical forces on cellular uptake
processes. Rotating magnetic fields of a few tens of mT have been used to increase
the uptake of iron-doped carbon nanotubes by cancer cells [102]. A benchtop
compact pulsed field system capable of producing fields of up to 10 T has been used
to deliver magnetic nanoparticles into cells that do not internalize nanoparticles by
phagocytosis [107].
1662 N. M. Dempsey

Magnetic Orientation
Homogeneous magnetic fields can apply a torque to orient macromolecules and
cells, which has potential applications in tissue engineering. The direction of
orientation depends on the molecule or cell type, due to anisotropy in susceptibility
(Fig. 13) [108]. A number of studies of magnetic orientation concerned the
polymerization of fibrous extracellular matrix proteins (fibrin, involved in blood
clotting, and collagen, found in connective tissue such as tendons, ligaments, skin,
etc.) in magnetic fields of the order of a few Tesla. While fibrin aligns parallel
to the field lines, collagen aligns perpendicular to the field lines. Magnetic fields
in the range of 1–8 T were also used to align red blood cells in gelatine. Normal
red blood cells, which have a donut-like shape, tend to align with their disk plane
parallel to the field. The degree of alignment depends on the field strength but is
independent of the oxidation state of the constituent hemoglobin, suggesting that

Fig. 13 Photographs of cells aligned either parallel or perpendicular to an applied magnetic field
[108]. (With kind permission from Elsevier)
34 Magnetism and Biology 1663

the torque acts on the diamagnetic cell membrane. In some people, polymerization
of deoxygenated hemoglobin deforms the shape of red blood cells from donut-
like to sickle-like and leads to cell rigidification. Blockage of these deformed cells
in small blood vessels leads to sickle cell anemia. When exposed to a magnetic
field, sickle cells align perpendicular to the field axis, owing to their anisotropic
paramagnetic susceptibility. Adherent cells such as osteoblasts that form bones,
endothelial cells that line the interior surface of blood vessels and lymphatic vessels,
smooth muscle cells, and Schwann cells, which are the principal nonneuronal cells
of the peripheral nervous system, have all been shown to align parallel to magnetic
field lines. The directional control of osteoblastic cells is studied because of its
potential use in bone fracture treatment [109]. Another application of magnetic
orientation in a homogeneous field concerns nerve regeneration [110]. Neurite
outgrowth from neurons was shown to be significantly greater on magnetically
aligned collagen gel than on nonaligned control gel and was found to depend on
the magnetic field strength. Such nerve regeneration could be exploited in tissue
engineering and regenerative medicine. In a complementary magnetic approach to
nerve regeneration, the field gradient force exerted by an inhomogeneous magnetic
field on magnetic nanoparticles attached to nerves has been used to induce tensile
forces which give rise to oriented neurite outgrowth [111] (Fig. 14).

Fig. 14 Schematic of oriented nerve regeneration mediated by functionalized magnetic nanopar-


ticles (f-MNP) [111]. (With kind permission from Elsevier)
1664 N. M. Dempsey

Magnetic Heating

Heat is known to have healing power, and whole-body or localized heating is used
in the treatment of a variety of medical conditions. An area of particular interest
is cancer treatment. Cancer cells stop growing at about 42 ◦ C, while normal cells
can tolerate higher temperatures. Heating to temperatures in the range of 42–45 ◦ C
is referred to as hyperthermia. When combined with irradiation or chemotherapy,
hyperthermia can efficiently damage cancer cells. Heating to temperatures above
50 ◦ C is referred to as thermoablation, and this leads to necrosis, coagulation, and
carbonization of tissue. The major challenge with cancer treatment by hyperthermia
is to restrict the temperature increase to the targeted region, i.e., the tumor. Localized
heating methods used in clinical settings for the treatment of superficial tumors
include ultrasound, microwave, and near-infrared irradiation. Alternating magnetic
fields can also be used to heat tissue, and most importantly deep tissue, through three
different mechanisms: dielectric loss in poorly conducting materials, eddy current
losses in conducting materials, and hysteresis losses in magnetic materials. The first
two mechanisms can give rise to direct heating of the tissue, while the third can lead
to highly localized heating of magnetic materials that then transfer heat through
conduction to the tissue. In 1957, Gilchrist used magnetic particle hyperthermia to
selectively heat the lymph nodes of dogs into which iron-oxide particles of size in
the range 20–100 nm had been injected [112]. Following this, ferromagnetic needles
were studied as thermal seeds, which could act as constant temperature seeds if
the material passes through its Curie temperature, or constant power seeds, if used
far below their Curie temperature [113, 114]. However, these fell out of favor as
they needed to be surgically implanted and aligned with the axis of the applied RF
field to be efficiently heated. A range of other magnetic powders were then studied
(Fe2 O3 /Fe3 O4 , Fe3 Co, Ni, ferrites, etc.), with by far the greatest number of studies
done on Fe2 O3 /Fe3 O4 nanoparticles.
Heating of magnetic nanoparticles exposed to an alternating magnetic field
occurs because of magnetization reversal and/or particle rotation in a fluid suspen-
sion. The physics behind magnetic particle hyperthermia is detailed in the chapter
of this handbook on magnetic nanoparticles. Here we recall that the most important
figure of merit is the specific absorption rate (SAR), which corresponds to the power
absorbed per mass of magnetic nanoparticles. While SAR depends on the intensity
and frequency of the magnetic field, the magnetic particle size and size distribution
have been found to be crucial parameters [115]. In vitro studies have been carried
out on cell cultures, to test the biocompatibility of the magnetic nanoparticles used.
Such studies serve to distinguish between thermal effects due to power absorption
from the alternating magnetic field and toxicity of the particles themselves.
When using an alternating magnetic field to produce heating through hysteresis
losses in magnetic materials within cancerous tissue, it is essential to minimize
heating due to eddy current losses in healthy tissue. Modeling is used to estimate
heating due to exposure to an alternating magnetic field. In the case of eddy
currents, the rate of heat production per unit volume of tissue of cylindrical shape is
given by.
34 Magnetism and Biology 1665

P = σ (π μo )2 (H0 f )2 r 2 (5)

where σ is the electrical conductivity of the tissue, μo is the permeability of vacuum,


Ho and f are the magnetic field amplitude and frequency, respectively, and r is the
distance from the cylinder axis [113]. To establish an upper limit to the value of H0 f
used in magnetic particle hyperthermia applied to a human thorax without inducing
eddy current heating, volunteers were exposed to an RF field produced by a single
turn induction coil surrounding their torso [113]. A field intensity of 35.8 Am−1
at a frequency of 13.56 MHz was found to be tolerable for extended periods. This
gave the tolerance limit of H0 f, known as the “Atkinson−Brezovich limit,” to be
4.85 × 108 A (m.s)−1 . Since the power varies as the square of the coil radius, the
alternating magnetic field limit imposed by eddy current heating increases as the
coil radius is reduced.
The primary challenge with using magnetic particle hyperthermia is to deliver
the particles to the targeted area. Direct injection into the tumor, or to blood vessels
supplying the tumor, is the most practical approach. Targeted delivery using either
magnetic nanoparticles labeled with tumor-specific antibodies or particle guidance
using inhomogeneous fields is also being studied. Another challenge is to have
a homogeneous distribution of magnetic nanoparticles within the target. Despite
human clinical trials of magnetic particle hyperthermia starting in 2004 [116], it is
not yet an approved clinical procedure. Much work is still going into optimizing the
magnetic nanoparticles and their modes of administration, as well as the magnetic
field systems and field parameter values used.

Effects of Magnetic Fields on Biological Organisms

In the preceding sections, we have described how magnetic fields are used in a range
of biomedical applications. It is important to ask what are the potential effects of
these fields on living organisms and what are the risks associated with using them.
This question is particularly relevant to MRI, as millions of patients are exposed
to high static magnetic fields, time-varying low-frequency gradient magnetic fields,
and high-frequency RF magnetic fields. Besides, thousands of healthcare workers
that operate and maintain MRI systems are repeatedly exposed to the stray static
fields produced by MRI systems.
Static magnetic fields may affect the human body through the Lorentz force felt
by moving charged particles in the bloodstream. Changes observed in electrocardio-
grams of animals in the presence and absence of a static magnetic field have been
attributed to electrical potentials induced by the Lorentz force. Electric fields and
currents can also be induced in bodies that are moving in a static magnetic field.
A static homogeneous field can produce a torque on objects such as paramagnetic
molecules showing anisotropic susceptibility and cause them to rotate, while a static
inhomogeneous field can displace magnetized objects through the field gradient
force. A number of studies have been carried out to assess the potential impact
1666 N. M. Dempsey

of static magnetic fields [110, 117]. For example, in vitro exposure of certain cells
to static fields (8 T during 3 h) was reported to result in increased levels of reactive
oxygen species, while genotoxic effects were reported in other cells exposed to 3 T
fields. Swelling and the formation of reactive oxygen species in cells positioned
close to arrays of permanent micromagnets were attributed to the very-high-gradient
static magnetic fields produced by such field sources (106 T/m at the surface of the
micromagnets) [118]. The potential effects of high-gradient static magnetic fields
on intracellular processes have been assessed from a theoretical point of view [119].
Following analysis of a large number of studies, the World Health Organization
(WHO) reported that both positive and negative effects were claimed on a range of
aspects studied (cell metabolic activity, cell membrane physiology, gene expression,
cell growth, genotoxicity) but that most data was not replicated. A WHO analysis
of literature related to the effects of static magnetic fields on animals concluded that
the movement of rodents in fields of 4 T or greater induced aversive responses and
conditioned avoidance related to discomfort. The observed effects were attributed
to the influence of the magnetic field on fluid in the inner ear. Some humans moving
in static fields produced by MRI machines reported experiencing vertigo, nausea,
poorer hand–eye coordination, and a metallic taste. The WHO concluded that there
is inadequate data to assess the health risks related to human exposure to static
magnetic fields.
Exposure to time-varying magnetic fields induces time-varying electric fields
and currents in the body. The intensities of the induced electric fields and currents
depend on the orientation of the time-varying magnetic field with respect to the
body, and they vary depending on the conductivity of the organs and tissue affected.
Induced currents are known to stimulate nerve and muscle cells, which can cause
discomfort. The onset of discomfort due to peripheral nerve stimulation is used to
set the upper limit for field exposure in terms of field amplitude and frequency.
Reports of people exposed to time-varying magnetic fields seeing flashes of light,
known as magnetophosphenes, are attributed to induced currents within the retina or
visual cortex. The International Agency for Research on Cancer (IARC) classified
the time-varying magnetic field associated with electricity transmission at 50–60 Hz
as possibly carcinogenic, based on epidemiological studies that reported increased
risk of childhood leukemia for magnetic field amplitudes greater than 0.4 μT.
However, there is no supporting evidence that such fields, referred to as extremely
low-frequency (EFL) fields, have an effect on cells or animals.
When a high-frequency RF electromagnetic field impacts upon tissue, the
induced electric field it produces in the tissue exerts forces on polarized molecules,
electrons, and ions, causing them to move, which leads to heating of the tissue.
This has to be considered when choosing field amplitude and frequency for
selective localized heating of magnetic nanoparticles during magnetic hyperthermia.
Exposure to RF fields may also influence the outcome of radical-pair biochemical
reactions within the body and could explain why magnetic navigation in birds is
disturbed by RF fields. The IARC classified RF fields as possibly carcinogenic to
humans, though evidence from experimental studies on animals and humans is very
limited.
34 Magnetism and Biology 1667

Transcranial Magnetic Stimulation (TMS)

Magnetic field pulses can be exploited to noninvasively stimulate neuronal and


muscular systems [110]. Magnetic stimulation of the brain is achieved with coils
producing fields in the range 1–2 T, while fields of 2–4 T are required to stimulate
the heart. The optimal pulse duration is of the order of 0.1–0.2 ms for the brain
and 1–2 ms for the heart. Transcranial magnetic stimulation (TMS) can be used to
induce a transient interruption of normal brain activity in a relatively restricted area
of the brain [110, 120] and is thus used in neurocognitive studies of brain function,
often together with electroencephalography, positron emission tomography, and
functional magnetic resonance imaging. It is also used in the diagnosis and treatment
of various neurological (depression, schizophrenia, post-traumatic stress disorder,
etc.) and movement disorders such as Parkinson’s disease. Figure-of-eight-shaped
coils in which current flows in opposite directions in adjacent circular coils allowed
higher induced electric field intensity and better field focusing compared to the
initially used single coils [121]. Modeling is used to estimate the focality of different
coil designs and has led to the development of coils with complex winding patterns,
examples of which are shown in Fig. 15 [122]. While figure-of-eight coils have been
shown to modulate brain activity to depths of 1.5–2.5 cm below the scalp, deep-
brain stimulation is achieved with more complex coils (3–4 cm with double-cone
coils, 4–6 cm with H-coils). However, greater penetration is achieved at the expense
of spatial resolution, resulting in the unwanted stimulation of tissue neighboring
the targeted region. The number of laboratories using TMS for studying the brain
and treating brain disorders has rapidly increased since it was first used by Barker
in 1985, and now it is applied in either single-pulse, paired-pulse, or repetitive-
pulse mode, depending on the application. While TMS is generally considered
safe, side effects, including transient headaches and neck pain, transient hearing
loss, temporary loss of consciousness, and, in very rare cases, seizures, have
been reported, mostly for high-frequency repetitive TMS. Safety and application
guidelines have been established for clinical practice and research [123]. Most of
the claims of therapeutic benefits of TMS in the treatment of various neurological,
psychiatric, and movement disorders await further support and evidence-based
clinical trials.

Recommended Magnetic Field Exposure Limits

The International Commission on Non-Ionizing Radiation Protection (ICNIRP) is


an independent organization, which provides scientific advice and guidance on
the health and environmental effects of nonionizing radiation. Their guidelines
related to human exposure to static magnetic fields are shown in Table 7 [124].
Occupational exposure to strong static magnetic fields concerns healthcare workers
using or servicing MRI systems and scientific and technical support staff working at
research facilities using strong magnetic fields. Lower static fields are experienced
1668

Fig. 15 Transcranial magnetic stimulation: Top, head models with different types of coils: double-cone coil (a), H-coil(b), halo-circular assembly coil, (c) and
figure-of-eight coil(d); bottom, corresponding simulated distribution of B-field (Tesla) with the contour outline of scalp and gray matter [122]. (Copyright:
© 2017 Lu, Ueno)
N. M. Dempsey
34 Magnetism and Biology 1669

Table 7 ICNIRP exposure guidelines for static magnetic fields


Exposure characteristics Magnetic flux density (T)
Occupational
Exposure of head and of trunk 2
Exposure of limbs 8
General public
Exposure of any part of the body 0.4

Fig. 16 ICNIRP exposure guidelines for low-frequency (1 Hz–100 kHz) magnetic fields [126].
(With kind permission from Wolters Kluwer Health, Inc.)

in industrial settings concerning electrolytic processes (e.g., chlorine or aluminum


production), permanent magnet, and magnetic material fabrication. Sources of
general public exposure to static fields include the Earth’s magnetic field, direct
current transmission lines, small permanent magnets used for fixation or latching,
and in transportation (magnetic levitation trains). A restriction of 0.5 mT for static
field exposure is recommended for members of the public with implanted electronic
medical devices and implants containing ferromagnetic material. Care also has to
be taken to avoid ferromagnetic objects such as tools or gas tanks being violently
attracted to field sources. The guidelines do not apply to the exposure of patients
undergoing medical diagnosis or treatment, and the ICNIRP has prepared a separate
statement concerning the protection of patients undergoing an MRI examination
[125].
The commission’s guidelines concerning exposure to low-frequency (1 Hz–
100 kHz) magnetic fields are plotted in Fig. 16 [126]. Restrictions are based on
established evidence regarding acute effects (e.g., peripheral and central nerve
stimulation, the induction of retinal phosphenes) and do not preclude interference
with, or effects on, medical devices such as metallic prostheses, cardiac pacemakers,
1670 N. M. Dempsey

and implanted defibrillators and cochlear implants. “Occupational exposure” refers


to adults exposed to time-varying magnetic fields at their place of work, generally
under known conditions. “General public” refers to individuals of all ages, and in
many cases, members of the public are unaware of their exposure to such fields.

Conclusions and Prospects

This chapter has given an overview of many, though certainly not all, links between
magnetism and biology. Advancing the study of magnetic fields produced by the
brain, through the improvement of existing and the development of new macro-
scopic and microscopic field detection tools and data analysis protocols, will help
us to progress in our understanding of how the brain works. This in turn should guide
the use of magnetic fields in the treatment of a range of brain disorders. The exploita-
tion of magnetic fields for other therapeutic applications requires validation through
systematic studies under very well-defined experimental conditions. A number of
biomedical applications involving the introduction of magnetic nanoparticles (e.g.,
magnetic particle-based imaging, hyperthermia, targeted drug delivery) or small
magnets (e.g., magnetically guided medical devices) into the human body are still at
the developmental stage and will require extensive animal and human clinical trials
before approval for clinical use. The further development of magnetic resonance
imaging in very high fields will lead to unprecedented resolution and access to
complex physiological processes, while at the low field end it will expand the use of
the technique to beyond hi-tech, high-cost environments. Developments in material
and device fabrication (e.g., microfabrication of magnetoelectric composites, 3D
printing of complex magnetic structures, etc.) should lead to advances in the
detection and exploitation of magnetic fields in biology and medicine. Fulfilling the
enormous potential at the interface between magnetism on one side and biology and
medicine on the other requires truly interdisciplinary research between physicists,
engineers, chemists, biologists, and medical doctors.

Further Reading

Magnetoreception and Magnetosomes in Bacteria, 2007, Edited by Dirk Schüler,


Springer
Biomagnetics: Principles and Applications of Biomagnetic Stimulation and Imag-
ing, 2017, Edited by Shoogo Ueno and Masaki Sekino, CRC Press
Magnetism in Medicine, A Handbook, 2007, Edited by Wilfried Andrä and Hannes
Nowak, Wiley-VCH
Clinical Applications of Magnetic Nanoparticles: From Fabrication to Clinical
Applications, 2018, Edited by Nguyen TK Thanh, CRC Press
34 Magnetism and Biology 1671

Acknowledgments The author is very grateful to Michael Winklhofer for very fruitful discussions
and feedback, to Kieron Ranno for artwork and to Mario Fratzl for assistance in putting the chapter
together.

References
1. Blakemore, R.P.: Magnetotactic bacteria. Annu. Rev. Microbiol. 36, 217–238 (1982)
2. Lefevre, C.T., Bazylinski, D.A.: Ecology, diversity, and evolution of magnetotactic bacteria.
Microbiol. Mol. Biol. Rev. 77, 497–526 (2013)
3. Bellini, S.: Su di un particolare comportamento di batteri d’acqua dolce. Institute of
Microbiology, University of Pavia, Pavia (1963)
4. Vargas, G., Cypriano, J., Correa, T., Leão, P., Bazylinski, D., Abreu, F.: Applications of
magnetotactic bacteria, magnetosomes and magnetosome crystals in biotechnology and
nanotechnology: mini-review. Molecules. 23, 2438–2425 (2018)
5. Bazylinski, D.A., Lefèvre, C.T., Schüler, D.: Magnetotactic bacteria. In: The Prokaryotes, pp.
453–494. Springer, Berlin/Heidelberg (2013)
6. Wiltschko, W., Wiltschko, R.: Magnetic orientation and magnetoreception in birds and other
animals. J. Comp. Physiol. A. 191, 675–693 (2005)
7. Johnsen, S., Lohmann, K.J.: The physics and neurobiology of magnetoreception. Nat. Rev.
Neurosci. 6, 703–712 (2005)
8. Wiltschko, W., Wiltschko, R.: Magnetic Compass of European Robins. Science. 176, 62–64
(1972)
9. Wiltschko, W., Munro, U., Ford, H., Wiltschko, R.: Red light disrupts magnetic orientation of
migratory birds. Nature. 364, 525–527 (1993)
10. Hanzlik, M., Heunemann, C., Holtkamp-Rötzler, E., Winklhofer, M., Petersen, N., Fleiss-
ner, G.: Superparamagnetic magnetite in the upper beak tissue of homing pigeons. Biometals.
13, 325–331 (2000)
11. Davila, A.F., Fleissner, G., Winklhofer, M., Petersen, N.: A new model for a magnetoreceptor
in homing pigeons based on interacting clusters of superparamagnetic magnetite. Phys. Chem.
Earth, Parts A/B/C. 28, 647–652 (2003)
12. Treiber, C.D., Salzer, M.C., Riegler, J., Edelman, N., Sugar, C., Breuss, M., Pichler, P.,
Cadiou, H., Saunders, M., Lythgoe, M., Shaw, J., Keays, D.A.: Clusters of iron-rich cells
in the upper beak of pigeons are macrophages not magnetosensitive neurons. Nature. 484,
367 (2012)
13. Schulten, K., Swenberg, C.E., Weller, A.: A biomagnetic sensory mechanism based on
magnetic field modulated coherent electron spin motion. Z. Phys. Chem. 111, 1–5 (1978)
14. Maeda, K., Henbest, K.B., Cintolesi, F., Kuprov, I., Rodgers, C.T., Liddell, P.A., Gust, D.,
Timmel, C.R., Hore, P.J.: Chemical compass model of avian magnetoreception. Nature. 453,
387–390 (2008)
15. Ahmad, M., Cashmore, A.R.: HY4 gene of A. thaliana encodes a protein with characteristics
of a blue-light photoreceptor. Nature. 366, 162–166 (1993)
16. Ritz, T., Adem, S., Schulten, K.: A model for photoreceptor-based magnetoreception in birds.
Biophys. J. 78, 707–718 (2000)
17. Hore, P.J., Mouritsen, H.: The radical-pair mechanism of magnetoreception. Annu. Rev.
Biophys. 45, 299–344 (2016)
18. Mouritsen, H.: Long-distance navigation and magnetoreception in migratory animals. Nature.
558, 50–59 (2018)
19. Caton, R.: The electric currents of the brain. Am. J. EEG Technol. 10, 12–14 (2015)
20. Berger, H.: Über das Elektrenkephalogramm des Menschen. Arch. Psychiatr. Nervenkr. 106,
165–187 (1937)
21. Baule, G., McFee, R.: Detection of the magnetic field of the heart. Am. Heart J. 66, 95–96
(1963)
1672 N. M. Dempsey

22. Cohen, D.: Magnetic fields around the torso: production by electrical activity of the human
heart. Science. 156, 652–654 (1967)
23. Cohen, D.: Magnetoencephalography: evidence of magnetic fields produced by alpha-rhythm
currents. Science. 161, 784–786 (1968)
24. Bison, G., Castagna, N., Hofer, A., Knowles, P., Schenker, J.L., Kasprzak, M., Saudan, H.,
Weis, A.: A room temperature 19-channel magnetic field mapping device for cardiac signals.
Appl. Phys. Lett. 95, 173701 (2009)
25. Pannetier-Lecoeur, M., Polovy, H., Sergeeva-Chollet, N., Cannies, G., Fermon, C., Parkko-
nen, L.: Magnetocardiography with GMR-based sensors. J. Phys. Conf. Ser. 303, 012054
(2011)
26. da Silva, F.L.: EEG and MEG: relevance to neuroscience. Neuron. 80, 1112–1128 (2013)
27. Roth, B.J., Wikswo Jr., J.P.: The magnetic field of a single axon. A comparison of theory and
experiment. Biophys. J. 48, 93–109 (1985)
28. Barry, J., Turner, M., Schloss, J.M., Glenn, D.R., Song, Y., Lukin, M., Park, H., Walsworth,
R.L.: Optical magnetic detection of single-neuron action potentials using quantum defects in
diamond. Proc. Natl. Acad. Sci. 113, 14133–14138 (2016)
29. Caruso, L., Wunderle, T., Lewis, C.M., Valadeiro, J., Trauchessec, V., Trejo Rosillo, J.,
Amaral, J.P., Ni, J., Jendritza, P., Fermon, C., Cardoso, S., Freitas, P.P., Fries, P., Pannetier-
Lecoeur, M.: In vivo magnetic recording of neuronal activity. Neuron. 95, 1283–1291
(2017)
30. Arami, H., Khandhar, A., Liggitt, D., Krishnan, K.M.: In vivo delivery, pharmacokinetics,
biodistribution and toxicity of iron oxide nanoparticles. Chem. Soc. Rev. 44, 8576–8607
(2015)
31. Pauling, L., Coryell, C.D.: The magnetic properties and structure of hemoglobin, oxyhe-
moglobin and carbonmonoxyhemoglobin. Proc. Natl. Acad. Sci. 22, 210–216 (1936)
32. Bren, K.L., Eisenberg, R., Gray, H.B.: Discovery of the magnetic behavior of hemoglobin: a
beginning of bioinorganic chemistry. Proc. Natl. Acad. Sci. 112, 13123–13127 (2015)
33. Moore, L.R., Fujioka, H., Williams, P.S., Chalmers, J.J., Grimberg, B., Zimmerman, P.A.,
Zborowski, M.: Hemoglobin degradation in malaria-infected erythrocytes determined from
live cell magnetophoresis. FASEB J. 20, 747–749 (2006)
34. Pankhurst, Q.A., Connolly, J., Jones, S.K., Dobson, J.: Topical review: applications of
magnetic nanoparticles in biomedicine. J. Phys. D. Appl. Phys. 36, R167–R181 (2003)
35. Wells, J., Kazakova, O., Posth, O., Steinhoff, U., Petronis, S., Bogart, L.K., Southern, P.,
Pankhurst, Q., Johansson, C.: Standardisation of magnetic nanoparticles in liquid suspension.
J. Phys. D: Appl. Phys. 50, 383003 (2017)
36. Levy, M., Luciani, N., Alloyeau, D., Elgrabli, D., Deveaux, V., Pechoux, C., Chat, S., Wang,
G., Vats, N., Gendron, F., Factor, C., Lotersztajn, S., Luciani, A., Wilhelm, C., Gazeau, F.:
Long term in vivo biotransformation of iron oxide nanoparticles. Biomaterials. 32, 3988–
3999 (2011)
37. Périgo, E.A., Hemery, G., Sandre, O., Ortega, D., Garaio, E., Plazaola, F., Teran, F.J.:
Fundamentals and advances in magnetic hyperthermia. Appl. Phys. Rev. 2, 041302 (2015)
38. Lauterbur, P.C.: Image formation by induced local interactions: examples employing nuclear
magnetic resonance. Nature. 242, 190–191 (1973)
39. Lauterbur, P.C.: Magnetic resonance zeugmatography. Pure Appl. Chem. 40, 149–157 (1974)
40. Mansfield, P., Grannell, P.K.: “Diffraction” and microscopy in solids and liquids by NMR.
Phys. Rev. B. 12, 3618–3634 (1975)
41. Arepally, A.: Targeted drug delivery under MRI guidance. J. Magn. Reson. Imaging. 27, 292–
298 (2008)
42. Ahrens, E.T., Bulte, J.W.M.: Tracking immune cells in vivo using magnetic resonance
imaging. Nat. Rev. Immunol. 13, 755–763 (2013)
43. Ogawa, S., Lee, T.M., Kay, A.R., Tank, D.W.: Brain magnetic resonance imaging with
contrast dependent on blood oxygenation. Proc. Natl. Acad. Sci. 87, 9868–9872 (1990)
44. Glover, G.H.: Overview of functional magnetic resonance imaging. Neurosurg. Clin. N. Am.
22, 133–139 (2011)
34 Magnetism and Biology 1673

45. Lee, I.-S., Preissl, H., Enck, P.: How to perform and interpret functional magnetic resonance
imaging studies in functional gastrointestinal disorders. J. Neurogastroenterol. Motil. 23,
197–207 (2017)
46. Silva, M.A., See, A.P., Essayed, W.I., Golby, A.J., Tie, Y.: Challenges and techniques for
presurgical brain mapping with functional MRI. NeuroImage: Clinical. 17, 794–803 (2018)
47. Cosmus, T.C., Parizh, M.: Advances in whole-body MRI magnets. IEEE Trans. Appl.
Supercond. 21, 2104–2109 (2011)
48. Vedrine, P., Aubert, G., Beaudet, F., Belorgey, J., Berriaud, C., Bredy, P., Donati, A.,
Dubois, O., Gilgrass, G., Juster, F.P., Meuris, C., Molinie, F., Nunio, F., Payn, A., Schild, T.,
Scola, L., Sinanna, A.: Iseult/INUMAC whole body 11.7 T MRI magnet status. IEEE Trans.
Appl. Supercond. 20, 696–701 (2010)
49. Budinger, T.F., Bird, M.D.: MRI and MRS of the human brain at magnetic fields of 14 T to 20
T: technical feasibility, safety, and neuroscience horizons. NeuroImage. 168, 509–531 (2018)
50. Ghazinoor, S., Crues III, J.V.: Low field MRI: a review of the literature and our experience in
upper extremity imaging. Clin. Sports Med. 25, 591–606 (2006)
51. Sarracanie, M., LaPierre, C.D., Salameh, N., Waddington, D.E.J., Witzel, T., Rosen, M.S.:
Low-cost high-performance MRI. Sci Rep. 5, 15177 (2015)
52. Gleich, B., Weizenecker, J.: Tomographic imaging using the nonlinear response of magnetic
particles. Nature. 435, 1214–1217 (2005)
53. Panagiotopoulos, N., Duschka, R.L., Ahlborg, M., Bringout, G., Debbeler, C., Graeser, M.,
Kaethner, C., Lüdtke-Buzug, K., Medimagh, H., Stelzner, J., Buzug, T.M., Barkhausen, J.,
Vogt, F.M., Haegele, J.: Magnetic particle imaging: current developments and future direc-
tions. Int. J. Nanomedicine. 10, 3097–3114 (2015)
54. Bakenecker, A., Ahlborg, M., Debbeler, C., Kaethner, C., Lüdtke-Buzug, K.: Magnetic
particle imaging. In: Precision Medicine, pp. 183–228. Elsevier (2018)
55. Bhat, V., Shenoy, K., Premkumar, P.: Magnets in dentistry. Arch. Med. Health Sci. 1, 73–79
(2013)
56. Sliker, L., Ciuti, G., Rentschler, M., Menciassi, A.: Magnetically driven medical devices: a
review. Expert Rev. Med. Devices. 12, 737–752 (2015)
57. Berry, M.V., Geim, A.K.: Of flying frogs and levitrons. Eur. J. Phys. 18, 307–313 (1997)
58. Kauffmann, P., Ith, A., O’Brien, D., Gaude, V., Boué, F., Combe, S., Bruckert, F., Schaack, B.,
Dempsey, N.M., Haguet, V., Reyne, G.: Diamagnetically trapped arrays of living cells above
micromagnets. Lab Chip. 11, 3153–3161 (2011)
59. Cugat, O., Delamare, J., Reyne, G.: Magnetic micro-actuators and systems (MAGMAS).
IEEE Trans. Magn. 39, 3607–3612 (2003)
60. Dempsey, N.M.: Hard magnetic materials for MEMS applications. In: Liu, J.P., Fullerton,
E., Gutfleisch, O., Sellmyer, D.J. (eds.) Nanoscale Magnetic Materials and Applications, pp.
661–684. Springer (2009)
61. Dempsey, N.M., Le Roy, D., Marelli-Mathevon, H., Shaw, G., Dias, A., Kramer, R.B.G.,
Viet Cuong, L., Kustov, M., Zanini, L.F., Villard, C., Hasselbach, K., Tomba, C., Dumas-
Bouchiat, F.: Micro-magnetic imprinting of high field gradient magnetic flux sources. Appl.
Phys. Lett. 104, 262401 (2014)
62. Walther, A., Marcoux, C., Desloges, B., Grechishkin, R., Givord, D., Dempsey, N.M.: Micro-
patterning of NdFeB and SmCo magnet films for integration into micro-electro-mechanical-
systems. J. Magn. Magn. Mater. 321, 590–594 (2009)
63. Dumas-Bouchiat, F., Zanini, L.F., Kustov, M., Dempsey, N.M., Grechishkin, R., Hassel-
bach, K., Orlianges, J.C., Champeaux, C., Catherinot, A., Givord, D.: Thermomagnetically
patterned micromagnets. Appl. Phys. Lett. 96, 102511 (2010)
64. Franzreb, M., Siemann-Herzberg, M., Hobley, T.J., Thomas, O.R.T.: Protein purification
using magnetic adsorbent particles. Appl. Microbiol. Biotechnol. 70, 505–516 (2006)
65. Ambashta, R.D., Sillanpää, M.: Water purification using magnetic assistance: a review. J.
Hazard. Mater. 180, 38–49 (2010)
66. Miltenyi, S., Müller, W., Weichel, W., Radbruch, A.: High gradient magnetic cell separation
with MACS. Cytometry. 11, 231–238 (1990)
1674 N. M. Dempsey

67. Herrmann, I.K., Bernabei, R.E., Urner, M., Grass, R.N., Beck-Schimmer, B., Stark, W.J.:
Device for continuous extracorporeal blood purification using target-specific metal nanomag-
nets. Nephrol. Dial. Transplant. 26, 2948–2954 (2011)
68. Frodsham, G., Pankhurst, Q.A.: Biomedical applications of high gradient magnetic sep-
aration: progress towards therapeutic haeomofiltration. Biomed. Eng./Biomed. Tech. 60,
393–404 (2015)
69. Stamopoulos, D., Benaki, D., Bouziotis, P., Zirogiannis, P.N.: In vitro utilization of ferromag-
netic nanoparticles in hemodialysis therapy. Nanotechnology. 18, 495102–495115 (2007)
70. Paul, F., Roath, S., Melville, D., Warhurst, D.C., Osisanya, J.O.S.: Separation of malaria-
infected erythrocytes from whole blood: use of a selective high-gradient magnetic separation
technique. Lancet. 318, 70–71 (1981)
71. Tietze, R., Zaloga, J., Unterweger, H., Lyer, S., Friedrich, R.P., Janko, C., Pöttler, M.,
Dürr, S., Alexiou, C.: Magnetic nanoparticle-based drug delivery for cancer therapy. Biochem.
Biophys. Res. Commun., 468, 463–470 (2015)
72. Pouponneau, P., Bringout, G., Martel, S.: Therapeutic magnetic microcarriers guided by
magnetic resonance navigation for enhanced liver chemoembilization: a design review. Ann.
Biomed. Eng. 42, 929–939 (2014)
73. Zakharchenko, A., Guz, N., Laradji, A.M., Katz, E., Minko, S.: Magnetic field remotely
controlled selective biocatalysis. Nature Catalysis, 1, 73–81 (2018)
74. Häfeli, U.O.: Magnetically modulated therapeutic systems. Int. J. Pharm. 277, 19–24 (2004)
75. Pamme, N.: Magnetism and microfluidics. Lab Chip. 6, 24–38 (2006)
76. Gijs, M.A.M., Lacharme, F., Lehmann, U.: Microfluidic applications of magnetic particles for
biological analysis and catalysis. Chem. Rev. 110, 1518–1563 (2010)
77. Hoshino, K., Huang, Y.-Y., Lane, N., Huebschman, M., Uhr, J.W., Frenkel, E.P., Zhang, X.:
Microchip-based immunomagnetic detection of circulating tumor cells. Lab Chip. 11, 3449–
3449 (2011)
78. Ahn, C.H., Allen, M.G., Trimmer, W., Jun, Y.N., Erramilli, S.: A fully integrated microma-
chined magnetic particle separator. J. Microelectromech. Syst. 5, 151–158 (1996)
79. Deng, T., Prentiss, M., Whitesides, G.M.: Fabrication of magnetic microfiltration systems
using soft lithography. Appl. Phys. Lett. 80, 461–463 (2002)
80. Osman, O., Toru, S., Dumas-Bouchiat, F., Dempsey, N.M., Haddour, N., Zanini, L.F.,
Buret, F., Reyne, G., Frénéa-Robin, M.: Microfluidic immunomagnetic cell separation using
integrated permanent micromagnets. Biomicrofluidics. 7, 054115–054112 (2013)
81. Pivetal, J., Toru, S., Frenea-Robin, M., Haddour, N., Cecillon, S., Dempsey, N.M., Dumas-
Bouchiat, F., Simonet, P.: Selective isolation of bacterial cells within a microfluidic device
using magnetic probe-based cell fishing. Sensors. Actuators. B. Chem. 195, 581–589 (2014)
82. Pamme, N., Manz, A.: On-Chip free-flow magnetophoresis: continuous flow separation of
magnetic particles and agglomerates. Anal. Chem. 76, 7250–7256 (2004)
83. Inglis, D.W., Riehn, R., Austin, R.H., Sturm, J.C.: Continuous microfluidic immunomagnetic
cell separation. Appl. Phys. Lett. 85, 5093–5095 (2004)
84. Cardoso, S., Leitao, D.C., Dias, T.M., Valadeiro, J., Silva, M.D., Chicharo, A., Silverio, V.,
Gaspar, J., Freitas, P.P.: Challenges and trends in magnetic sensor integration with microflu-
idics for biomedical applications. J. Phys. D. Appl. Phys. 50, 213001 (2017)
85. Lim, B., Vavassori, P., Sooryakumar, R., Kim, C.: Nano/micro-scale magnetophoretic devices
for biomedical applications. J. Phys. D: Appl. Phys, 50, 033002 (2017)
86. Crick, F.H.C., Hughes, A.F.W.: The physical properties of cytoplasm. Exp. Cell Res. 1, 37–80
(1950)
87. Strick, T.R., Allemand, J.F., Bensimon, D., Bensimon, A., Croquette, V.: The elasticity of a
single supercoiled DNA molecule. Science. 271, 1835–1837 (1996)
88. Gosse, C., Croquette, V.: Magnetic tweezers: micromanipulation and force measurement at
the molecular level. Biophys. J. 82, 3314–3329 (2002)
89. van Oene, M.M., Dickinson, L.E., Pedaci, F., Köber, M., Dulin, D., Lipfert, J., Dekker, N.H.:
Biological magnetometry: torque on superparamagnetic beads in magnetic fields. Phys. Rev.
Lett. 114, 218301 (2015)
34 Magnetism and Biology 1675

90. Danilowicz, C., Greenfield, D., Prentiss, M.: Dissociation of ligand−receptor complexes
using magnetic tweezers. Anal. Chem. 77, 3023–3028 (2005)
91. Kilinc, D., Lee, G.U.: Advances in magnetic tweezers for single molecule and cell biophysics.
Integr. Biol. 6, 27–34 (2014)
92. Wang, N., Butler, J., Ingber, D.: Mechanotransduction across the cell surface and through the
cytoskeleton. Science. 260, 1124–1127 (1993)
93. Sniadecki, N.J., Anguelouch, A., Yang, M.T., Lamb, C.M., Liu, Z., Kirschner, S.B., Liu, Y.,
Reich, D.H., Chen, C.S.: Magnetic microposts as an approach to apply forces to living cells.
Proc. Natl. Acad. Sci. 104, 14553–14558 (2007)
94. Bidan, C.M., Fratzl, M., Coullomb, A., Moreau, P., Lombard, A.H., Wang, I., Balland, M.,
Boudou, T., Dempsey, N.M., Devillers, T., Dupont, A.: Magneto-active substrates for local
mechanical stimulation of living cells. Sci. Rep. 8, 1464 (2018)
95. Dobson, J.: Remote control of cellular behaviour with magnetic nanoparticles. Nat. Nanotech-
nol. 3, 139–143 (2008)
96. Brunet, T., Bouclet, A., Ahmadi, P., Mitrossilis, D., Driquez, B., Brunet, A.-C., Henry, L.,
Serman, F., Bealle, G., Menager, C., Dumas-Bouchiat, F., Givord, D., Yanicostas, C., Le
Roy, D., Dempsey, N.M., Plessis, A., Farge, E.: Evolutionary conservation of early mesoderm
specification by mechanotransduction in Bilateria. Nat. Commun. 4, 2821 (2013)
97. Mitrossilis, D., Röper, J.-C., Le Roy, D., Driquez, B., Michel, A., Ménager, C., Shaw, G., Le
Denmat, S., Ranno, L., Dumas-Bouchiat, F., Dempsey, N.M., Farge, E.: Mechanotransductive
cascade of Myo-II-dependent mesoderm and endoderm invaginations in embryo gastrulation.
Nat. Commun. 8, 13883 (2017)
98. Ogiue-Ikeda, M., Sato, Y., Ueno, S.: A new method to destruct targeted cells using
magnetizable beads and pulsed magnetic force. IEEE Trans. Nanobiosci. 2, 262–265 (2003)
99. Kim, D.-H., Rozhkova, E.A., Ulasov, I.V., Bader, S.D., Rajh, T., Lesniak, M.S., Novosad, V.:
Biofunctionalized magnetic-vortex microdiscs for targeted cancer-cell destruction. Nat.
Mater. 9, 165–171 (2010)
100. Mansell, R., Vemulkar, T., Petit, D.C.M.C., Cheng, Y., Murphy, J., Lesniak, M.S., Cow-
burn, R.P.: Magnetic particles with perpendicular anisotropy for mechanical cancer cell
destruction. Sci. Rep. 7, 4257 (2017)
101. Wang, B., Bienvenu, C., Mendez-Garza, J., Lançon, P., Madeira, A., Vierling, P., Di
Giorgio, C., Bossis, G.: Necrosis of HepG2 cancer cells induced by the vibration of magnetic
particles. J. Magn. Magn. Mater. 344, 193–201 (2013)
102. Liu, D., Wang, L., Wang, Z., Cuschieri, A.: Magnetoporation and magnetolysis of cancer
cells via carbon nanotubes induced by rotating magnetic fields. Nano Lett. 12, 5117–5121
(2012)
103. Muroski, M.E., Morshed, R.A., Cheng, Y., Vemulkar, T., Mansell, R., Han, Y., Zhang, L.,
Aboody, K.S., Cowburn, R.P., Lesniak, M.S.: Controlled payload release by magnetic field
triggered neural stem cell destruction for malignant glioma treatment. PLoS One. 11,
e0145129 (2016)
104. Cheng, Y., Muroski, M.E., Petit, D.C.M.C., Mansell, R., Vemulkar, T., Morshed, R.A., Han,
Y., Balyasnikova, I.V., Horbinski, C.M., Huang, X., Zhang, L., Cowburn, R.P., Lesniak, M.S.:
Rotating magnetic field induced oscillation of magnetic particles for in vivo mechanical
destruction of malignant glioma. J. Control. Release. 223, 75–84 (2016)
105. Plank, C., Zelphati, O., Mykhaylyk, O.: Magnetically enhanced nucleic acid delivery. Ten
years of magnetofection—progress and prospects. Adv. Drug. Deliv. Rev. 63, 1300–1331
(2011)
106. Xiaofan, D., Jing, W., Quan, Z., Luwei, Z., Sijia, W., Zhenxi, Z., Cuiping, Y.: Advanced
physical techniques for gene delivery based on membrane perforation. Drug Deliv. 25, 1516–
1525 (2018)
107. Uzhytchak, M., Lynnyk, A., Zablotskii, V., Dempsey, N.M., Dias, A.L., Bonfim, M., Lunova,
M., Jirsa, M., Kubinová, Š., Lunov, O., Dejneka, A.: The use of pulsed magnetic fields to
increase the uptake of iron oxide nanoparticles by living cells. Appl. Phys. Lett. 111, 243703–
243706 (2017)
1676 N. M. Dempsey

108. Ueno, S., Sekino, M.: Biomagnetics and bioimaging for medical applications. J. Magn. Magn.
Mater. 304, 122–127 (2006)
109. Kotani, H., Kawaguchi, H., Shimoaka, T., Iwasaka, M., Ueno, S., Ozawa, H., Nakamura, K.,
Hoshi, K.: Strong static magnetic field stimulates bone formation to a definite orientation in
vitro and in vivo. J. Bone Miner. Res. 17, 1814–1821 (2002)
110. Ueno, S., Sekino, M.: Biomagnetics: Principles and Applications of Biomagnetic Stimulation
and Imaging. CRC Press, Boca Raton (2016)
111. Riggio, C., Calatayud, M.P., Giannaccini, M., Sanz, B., Torres, T.E., Fernández-Pacheco, R.,
Ripoli, A., Ibarra, M.R., Dente, L., Cuschieri, A., Goya, G.F., Raffa, V.: The orientation of
the neuronal growth process can be directed via magnetic nanoparticles under an applied
magnetic field. Nanomed. Nanotechnol. Biol. Med. 10, 1549–1558 (2014)
112. Gilchrist, R.K., Medal, R., Shorey, W.D., Hanselman, R.C., Parrott, J.C., Taylor, C.B.:
Selective inductive heating of lymph nodes. Ann. Surg. 146, 596–606 (1957)
113. Atkinson, W.J., Brezovich, I.A., Chakraborty, D.P.: Usable frequencies in hyperthermia with
thermal seeds. IEEE Trans. Biomed. Eng. BME-31, 70–75 (1984)
114. Brezovich, I.A., Atkinson, W.J., Lilly, M.B.: Local hyperthermia with interstitial techniques.
Cancer Res. 44, 4752s–4756s (1984)
115. Jordan, A.: Hyperthermia classic commentary: “Inductive heating of ferrimagnetic particles
and magnetic fluids: Physical evaluation of their potential for hyperthermia” by Andreas
Jordan et al., International Journal of Hyperthermia, 1993;9:51-68. Int J Hyperthermia. 25,
512–516 (2009)
116. Gneveckow, U., Jordan, A., Scholz, R., Brüß, V., Waldöfner, N., Ricke, J., Feussner, A., Hilde-
brandt, B., Rau, B., Wust, P.: Description and characterization of the novel hyperthermia- and
thermoablation-system MFH® 300F for clinical magnetic fluid hyperthermia. Med. Phys. 31,
1444–1451 (2004)
117. Schenck, J.F.: Physical interactions of static magnetic fields with living tissues. Prog. Biophys.
Mol. Biol. 87, 185–204 (2005)
118. Zablotskii, V., Syrovets, T., Schmidt, Z.W., Dejneka, A., Simmet, T.: Modulation of mono-
cytic leukemia cell function and survival by high gradient magnetic fields and mathematical
modeling studies. Biomaterials. 35, 3164–3171 (2014)
119. Zablotskii, V., Polyakova, T., Lunov, O., Dejneka, A.: How a high-gradient magnetic field
could affect cell life. Sci. Rep., 6, 37407 (2016)
120. Barker, A.T., Jalinous, R., Freeston, I.L.: Non-invasive magnetic stimulation of human motor
cortex. Lancet. 325, 1106–1107 (1985)
121. Ueno, S., Tashiro, T., Harada, K.: Localized stimulation of neural tissues in the brain by
means of a paired configuration of time-varying magnetic fields. J. Appl. Phys. 64, 5862–
5864 (1988)
122. Lu, M., Ueno, S.: Comparison of the induced fields using different coil configurations during
deep transcranial magnetic stimulation. PLoS One. 12, e0178422 (2017)
123. Rossi, S., Hallett, M., Rossini, P.M., Pascual-Leone, A., Group, T.S.O.T.C: Safety, ethical
considerations, and application guidelines for the use of transcranial magnetic stimulation in
clinical practice and research. Clin. Neurophysiol. 120, 2008–2039 (2009)
124. International Commission on Non-Ionizing Radiation Protection: Guidelines on limits of
exposure to static magnetic fields. Health. Phys. 96, 504–514 (2009)
125. Medical magnetic resonance (MR) procedures: protection of patients. Health Phys. 87, 197–
216 (2004)
126. International Commission on Non-Ionizing Radiation Protection: Guidelines for limiting
exposure to time-varying electric and magnetic fields (1 Hz to 100 kHz). Health Phys. 99,
818–836 (2010)
34 Magnetism and Biology 1677

Nora Dempsey received her PhD from Trinity College Dublin,


Ireland, in 1998. Since then she has been based at Institut Néel,
CNRS Grenoble in France. She works on the study and devel-
opment of functional magnetic materials in film form, with an
emphasis on hard magnetic materials for energy and bio-related
applications.
Index

A Alnico(s), 698, 702–704, 1390, 1391,


Aberration correction, 1213 1412, 1421
Ab-initio calculations, 220, 516, 517, 575 magnets, 28
Ab-initio theory, 189 α−waves, 1639, 1641
Ablation process, 1182 Alternating gradient technique, 1139
Abrikosov flux, 557 Aluminium-nickel-cobalt alloys
Abrikosov-Shul resonance, 934 (AlNiCo), 1086
Achavalite, 894 Aluminosilicates, 848
Acmite, 898, 901 Amakinite, 869
AC-susceptibility measurements, 1134 Ammonium hydroxide, 1020
Actinide anisotropy, see Anisotropy Amorphous (FeCo)B, 581
Actinide compounds, 890 Co-based alloy, 1470
Action integral, 1100 Fe-based alloy, 1468
Activation volume, 1386 ferromagnetic alloys, 35
Actuators, 1421 magnets, 165
Additive manufacturing, 1400, 1401 rare-earth intermetallic
Adherent cells, 1663 alloys, 36
Adiabatic approximation, 233, 240 Ampere’s law, 356, 1594
Adiabatic demagnetisation (ADR) materials, Ampère, André-Marie, 10
1493, 1495, 1496 Anderson Hamiltonian, 931
Adiabatic dynamics, 204–208 Anderson impurity model (AIM), 195
Adiabatic temperature change, 1507 Anderson-Mott localization, 945
Advanced functional recycling, 1408 Anderson-Mott transition, 934, 943
Aerotaxis, 1636 Andradite, 871, 898
Aether, 16 Angle-resolved photoemission spectroscopy
Aggregation, 1156 (ARPES), 1237
Airgap, 1412 Angular frequency, 1314
Akaganénite, 868 Angular momentum, 21, 486, 512, 982,
Algebraic eigenvalue problem, 196 1299, 1300
All-optical switching of magnetization, 521, Animal magnetism, 8
525–527 Anisometer, magnetic, 1137
ferrimagnetic alloys, 521–523 Anisotropic exchange, 1055
ferromagnetic systems, 523–525 Anisotropic HDDR, 1400
helicity-dependent process, 521 Anisotropic magnetoresistance (AMR), 18,
Alloy(s), 33 39, 44, 45, 440, 442–444, 446,
anisotropies, 146 447, 1117, 1118, 1349, 1535,
films, 1191, 1192 1536, 1570
Almandine, 871, 898 sensors, 44, 1118
Almeida-Thouless line, 956 Anisotropic resonance, 1306

© Springer Nature Switzerland AG 2021 1679


J. M. D. Coey, S. S. P. Parkin (eds.), Handbook of Magnetism and Magnetic
Materials, https://doi.org/10.1007/978-3-030-63210-6
1680 Index

Anisotropic Zeeman interaction, 1306–1308 metallic elements, 810


Anisotropy, 144, 823, 1051, 1054, 1307 memory, 1570, 1571
anisotropy constants, 139–179 moments, 1570
anisotropy field, 114, 166 spin cycloid, 605
Ba and Sr ferrites, 156 superexchange, 948
constants, 27, 105, 107–110, 112–116, Antiferromagnetic order,
139–142, 144, 145, 147, 148, superconductivity, 627
153, 156, 160, 169, 1137, 1374, microscopic consideration, 643–648
1386, 1390 phenomenological description, 640–642
and crystal structure, 109–111 Antiferromagnetism (AFM), 23, 54, 56, 62–64,
cubic (esp. Fe and Ni), 109, 115, 78, 84, 86, 87, 92–94, 96, 97, 165
166, 170 interatomic exchange and magnetic
4d, 5d, and 5f anisotropies, 149 order, 55
easy axis, 105, 165 weak ferromagnets, 165
energy, 400 Antiphase boundary, 1019
energy per rare-earth atom, 139 Antiresonance frequency, 1320
4f anisotropy, 132, 144 Apoptosis, 1660
field, 27, 115, 365–366, 1392, 1394, 1398 Applications of MBE
Gd anisotropy contribution, 105 Dirac and Weyl semimetals, 1173, 1175
hexagonal, 113 HTC, 1175, 1176
higher-order, 114–115 semiconductors, 1169
iron-type anisotropy, 109 three-temperature-method, 1169
of magnetic phenomena, 1393 TI, 1171, 1172
magnetocrystalline, 1048, 1059, 1068 transition metal dichakcogenides
magnetoelastic anisotropy, 166, 167 1175–1177
nanostructures, 168 two-dimensional magnetic materials,
noncubic crystal structures, 110 1169–1171
Néel anisotropy, 162 Aqueous syntheses, 1020
perpendicular magnetic, 1055 Arbitrary waveform generator (AWG), 1326
point groups and space groups, 111, 176 Archaeomagnetism, 37
random, 116, 169–171 Artificial multiferroics, 371
shape anisotropy, 105, 168 Artificial permanent magnets, 1390
shape of rare-earth ions, 133 Artificial spin-ice structures, 1061, 1063
single atom, 1070 Artificially-engineered magnetic materials
spin chain, 153 artificial atoms and Coulomb blockade,
temperature dependence, 141, 144 1064, 1068
tetragonal, 112, 113 lithographically-patterned 2D
trigonal, 113 ferromagnetic arrays,
two-ion anisotropy, 144, 159, 161–164 1060, 1063
typical anisotropy constants, 115 nanocomposite materials, 1048
uniaxial vs. biaxial, 105, 116, 169 nanowires, 1058, 1060
See also 4f anisotropyrare-earth single atom manipulation and
anisotropy, 133 measurements, 1068, 1071
Anomalous Hall effect, 18, 47, 188, 205, 458 thin-film structures, 1049, 1058
Anomalous loss, 1437 Asymmetric Bloch walls, 418
Anti-damping magnetic switching, Asymmetric Néel wall, 419
1339–1342 Atkinson–Brezovich limit, 1665
Anti-magnetic stainless steel, 666 Atom kinetic energy, 1192
Anticlastic curvature, 573 Atomic crystal-field charges, 107
Antiferromagnet (AFM), 207, 276, 1049, Atomic flux/surface temperature, 1155
1051, 1273 Atomic force microscopy, 1218
anisotropy, 165 Atomic layer deposition (ALD), 908–909
coupling, 928 Atomic level theory, 392
ground state, 1505 Atomic sphere approximation (ASA), 234
Index 1681

Atomic-scale magnetism, 161 Biomagnetism, see Magnetic navigation


Atomic-scale random-anisotropy effects, 170 Biomarkers, 1023
Atomistic LLG (ALLG) models, 516 Biomedical applications, 1020, 1021
Auger electron spectroscopy (AES), 1168 magnetic nanoparticles, 1644–1646
Augmented spherical wave (ASW), 234 Biomedical sensing applications, 1023
Austenite, 1390 Biot, Jean-Baptiste, 10
Automatic frequency control (AFC), 1323 Biot-Savart law, 1087
Automotive sensor, 1535 Biotargets, 1653
Avalanche photo diode (APD), 1233 Bitter coil, 1653
Avian magnetic navigation Bitter method, 1029
magnetite-based magnetoreception, 1637 Bitter-type magnets, 1097
radical-pair based magnetoreception, Bixbyite, 877, 889
1637–1639 Bloch, Felix, 282
Bloch-Bloemergen equations, 1027
Bloch equations, 1239, 1303, 1304
B Bloch lines, 418
Ba2 Fe12 O19 , 1653 Bloch sphere, 1001
Back-shearing, 1412 Bloch’s law, 84, 93, 94
Backward extrusion, 1388, 1399 Bloch states, 196–198
Balanced photodiode configuration, 490 Bloch wall, 415–418, 422, 423, 429
Balanced photodiode scheme, 491 parameter, 375
Ballistic molecular beams, 1165 Bloch wave functions, 197, 939
B20 alloys, 89 Bloch-wall width, 1380, 1386
Band Jahn–Teller effect, 724 Blocking temperature, 984
Band structure(s), 662, 663 Bloembergen-Rowland mechanism, 949
methods, 196–201 Blood-Oxygen-Level-Dependent (BOLD)
Band-filling dependence, 829 contrast, 1647
Band-filling effect, 150 Bogoliubov transformation, 94
Barbier plot, 1387 Bohm-Aharonov oscillations, 992
Barium ferrite, 1392 Bohr, Niels, 21
Barocaloric effect (BCE), 766, 768 Bohr magneton, 21, 350, 351, 983, 1334, 1338,
Beam splitter (BS), 499 1342, 1594, 1597
Bernouilli, Daniel, 9 Bohr radius, 944
Berry curvature, 204, 205 Bohr-van Leeuwen theorem, 106, 188
Berry phase, 98, 106, 459 Boltzmann equations, 515
interference, 992 Boltzmann formalism, 201
Bertotti’s model, 1445–1448 Boltzmann’s constant, 354, 1386, 1597
Bethe ansatz, 63, 93 Born approximation, 941
Bethe-Slater-Néel curve, 75–76 Bornite, 893
Bi2 Sr2 CaCu2 O8 , 559 Bound magnetic polarons (BMPs), 942–944
Biaxial strain, 904 Bragg scattering, 1257, 1259, 1260, 1266
Bicycle dynamo, 1423 Brain waves, 1641
Big data revolution, 39 Branched domains, 1380
Binary Co-Sm equilibrium phase diagram, 758 Bravais lattice vector, 233
Binodal decomposition, 936 Breit interaction, 227
Biocompatibility, 1644 Brillouin functions, 84, 85
Biodegradation, 1645, 1646 Brillouin light scattering spectroscopy (BLS),
Biodistribution, 1644 288, 571, 1350
Biological separation, 1621 Brinkman-Rice phase, 66
Biomagnetic fields, 1639, 1641 Brown’s micromagnetic equation, 376–377
Biomagnetic measurements, 1532 Euler method, 377–379
Biomagnetism Ritz method, 380–381
iron, 1642, 1643 Brown’s paradox, 115, 1386
magnetic fields, 1639–1642 Brown, William Fuller, 26
1682 Index

Brownian rotation, 1024 Charge-transfer insulator, 66, 860


Brushed DC motor, 1422 Charge transfer state, 932
Bubble memory, 880 Chemical-mechanical polishing (CMP), 1160
Buffer layer, 1159–1164 Chemical potential effects, 1068
Chemical vapor deposition (CVD), 908, 1461
C Chemotherapy, 1024
Cahn-Hillard equation, 936 Chimneys, 1189, 1190
Calibration and metrology, 1141–1143 Chirality, 1608
Cancer cells, 1660 Chirality-induced Hall effect, 208
Cancerous tumors, 1024 Chromite, 870
Canted antiferromagnetic spins, 605 Chromium, 812
Cantilever magnetometers, 1137–1138 crystal structure, 668
Carbon monoxide (Co), 148, 164, 166 magnetic structure, 669
Carbon steel, 1452 Circuit, magnetic, 1091
Carbonates, 903 Circular dichroism, 116
Cardiovascular imaging, 1647 Circularly polarized eigen modes, 483
Carrier-mediated spin-spin coupling Circularly polarized laser pulses, 521
p − d Zener model, 949–950 Circularly polarized light, 478
Bloembergen-Rowland mechanism, 949 Classical electrodynamics, 12
Curie temperature, 950–952 Clausius-Mosotti formula, 1481
RKKY interaction, 949 Clebsch-Gordan coefficients, 68
Carrollite, 894 Clebsch-Gordon coefficients, 1230
Cartoon-like “shaking-ghost” Closed loop measurement, 1544
interpretation, 140 Closed-circuit magnetometry, 1135–1136
Cation(s), 852, 853, 860, 861, 863, 866, Co, 86, 1644, 1654
867, 869, 875–877, 879–881, 884, Co complexes (high spin vs. low spin), 69
887, 889, 894, 897–899, 901, 902, Co-doping, 966
911, 912 Co-ferrite, 1392
distribution, 870 Co/GaAs (001), 819–824
Cation-substitutional randomly distributed Cox Mn1−x /GaAs (001), 829–832
Mn ions, 964 Coarse-grained powder, 1400
Cattierite, 893 Cobalt ferrite, 870
Cell boundaries, 1395 Cobalt(II) oxide (CoO), 84, 166
Cell destruction, 1660, 1661 Coefficient of refrigerant performance
Cellular scale, 1025 (CRP), 1506
Cellulose fibers, 1022 Coercive field, 373, 1385
Cerebral Blood Flow (CBF), 1647 Coercivity, 27, 1385, 1387, 1394, 1398, 1403
Certified reference materials (CRM), Coherent potential approximation (CPA), 199,
1142, 1143 201, 213, 247, 248
cgs units, 48 Coherent x-ray imaging (CXI), 1235–1237
Chalcocite, 893 Coil
Chalcogenides, 892–895 Helmholtz, 1090
Chalcopyrite, 893 induction, 1110, 1111, 1113
Chandrasekhar-Clogston limit, 629 pick-up, 1111
Channel resistivities, 445 sensors, 1530, 1531
Characteristic length scales, 372 single-turn, 1102–1103
critical diameter for uniform rotation, 373 Cole exponents, 1451
exchange length, 372 Colloid, 1062
mesh size, in micromagnetic Colossal magnetoresistance (CMR), 31,
simulations, 376 448–451, 883, 887, 945
wall parameter, 375–376 Colossal magnetoresistive (CMR), 1282
Charge densities, 611 Columnar growth, 1158
Charge-density wave (CDW), 669, 1273 Compass, 5–7
Charge order, 856, 935 needles, 1388
Index 1683

Compensated ferrimagnetism (CFiM), 86 Coulomb repulsion, 644


Compensation point, 1480 Coulson-Fischer wave functions, 62, 66, 78
Complementary metal–oxide–semiconductor Coulumb blockade, 999
(CMOS) technology, 1554 Cr, 87
Complex Kerr rotation, 482, 483, 489 Cr/GaAs (001), 812, 813
Composite magnetoelectrics, 608 Cr/GaAs (001). LICENSE #: 4812830486629,
direct magnetoelectric effect, 609 813
integrating magnetic and piezoelectric CrBr3 , 56
materials, 607 Creep induced anisotropy, 400
inverse magnetoelectic effect, 607 Critical competing technologies, 1408
inverse magnetoelectric effect, 609–613 Critical exponents, 34
Composite multiferroics Critical material, 1408
electric field switching, 602 Critical single-domain-particle diameter, 1380
electric-field control of magnetism, 608 Critical slowing down, 75, 95
magnetic and ferroelectric materials, 607 Crocidolite, 898, 900
magnetic and plain-dielectric, 607 Cronstedtite, 898, 899
Compression molding, 1400 Cross-tie wall, 416, 418
Compression molding with subsequently Crossed polarizer configuration, 490, 499
sintering (CMS), 703 Crustal abundances, 848, 850
Computer simulations, 348 Cryo-transmission electron microscopy,
Concentration profiles, 1398 1028
Condensation energy, 628 Crystal anisotropy energy, 364
Condensed magnetic semiconductors (CMS), anisotropy field, 365–366
931, 966 cubic anisotropy, 364–365
Conductive atomic force microscopy uniaxial anisotropy, 365
(CAFM), 1036 Crystal elasticity, 552
Conductivity tensor, 436 Crystal field, 576, 853–855, 858
Configuration interactions, 63 effects, 487, 511
Connolly-Williams method, 221, 223 expansion, 122–123
Conservation of angular momentum, interaction, 27, 107, 118, 123, 125, 132,
510–512 138, 150, 155, 156, 162, 171, 174
Continuous wave (CW), 1303, 1321 parameters, 122, 123, 140
Contrast agents, 1239, 1647, 1649, 1650 stabilization energy, 855
Conventional GOSS, 1457, 1458 theory, 931
Cooper-pairing, 627, 644, 645 Crystal-field splitting
Coplanar waveguide (CPW) 3d electrons, 120
configuration, 1324 cubic e-t2 crystal-field splitting, 119
Copper ferrite, 870 Crystal-field theory, 116–131
CoPt, 86 origin, 117
Corbino disc, 1115, 1116 one-electron crystal-field splitting,
Core, magnetic, 1090–1091 118–121
Core-shell NPs, 1032, 1033 Crystalline electric field (CEF), 228, 229,
Correlations, 62, 65, 68, 78, 94 1374, 1394, 1404
energy, 62 Crystalline environments, 119
hole, 55 Crystalline system, 1314
Corrosion, 1392, 1397, 1601, 1610 Crystal structure, 109, 897
Cotunnelling, 1066, 1067 Cubanite, 893
Coulomb blockade (CB), 1064, 1065 Cubic anisotropy, 141, 364–365, 1118
Coulomb diamonds, 1065 Cubic crystals, 113
Coulomb energy, 931 Cubic magnetocrystalline
Coulomb gap, 935, 1066 anisotropy, 829
Coulomb interaction, 55, 56, 62, 73, 351, 858, Curie, Pierre, 19
937, 1374 Curie constant, 1597
Coulomb potential, 190, 941 Curie law, 1596
1684 Index

Curie temperature, 23, 33, 34, 69, 83–84, Diffuse magnetic scattering, 1276
92, 130, 223, 354, 439, 443, 448, Diffusion, 1603
552, 627, 648, 929, 950–952, equation, 1442
961–963, 1392 layer, 1602
Curie-Weiss law, 20, 247, 687 Difluorides, 890
Current sensing, 1545 Dilute ferromagnetic semiconductors (DFSs),
Current-in-plane (CIP), 446, 449 928–929, 957–960
Current-perpendicular-to-plane (CPP), 446, Dilute magnetic materials (DMMs), 928
449, 465 band and localized spins, exchange
Curvature measurement, 572 interactions, 936–938
Cycloidal spin spiral, 601, 602, 604 dilute ferromagnetic insulators and
Cyclotron motion, 1127 topological insulators, 960–963
dilute magnetic semiconductors,
929–930
D dilute magnetic topological materials, 930
Damon-Eshbach spin-wave mode, 283, 294 dominant spin–spin interactions,
DC magnets, 1094 946–952
high-field resistive magnets, energy states of magnetic dopants, in
1095–1098 solids, 931–936
hybrid magnets, 1098 heterogenous magnetic semiconductors and
DC servomotor, 1422 oxides, 930, 964–967
Dc sputtering, 907 p-type dilute ferromagnetic
de Gennes factor, 69, 85, 86, 130 semiconductors, 928–929,
de-Gennes polymer analogy, 92 957–960
Dead layers, 910–911 sp-d( f ) exchange interactions, effects of,
Debye model, 1450 938–946
Decoherence, 265–268, 274, 275 spin-glass systems, 952–957
Decomposition, 1156 Dilute magnetic oxides, 911–914
Decoration methods, 1140 Dilute magnetic semiconductors (DMSs), 221,
Degenerate Fermi-surface crossing 912, 929–930, 938, 939, 945–947,
(DFSC), 152 953
3d elements, 1394 with rare earth metals, 948
Demagnetization factor, 15, 1377 with transition metals, 948
Demagnetization tensor, 1377, 1390 Dilute magnetic topological materials, 930
Demagnetization time, 506 Dimensionless numbers
Demagnetizing field, 14, 15, 358–361, 401, fluidic, 1603
1376, 1377 MHD, 1606
Demand for Nd-Fe-B, 1409 Dioxides, 877–878
Dense stripe domains, 412 Dipolar interaction, 1314
Density functional theory (DFT), 78, 79, 81, Dipolar nets, 268–271
95, 154, 155, 190, 574, 639 Dipole fields, 358–359
Density of states (DOS), 73, 76, 77, 193, Dipole selection rule, 485
195, 196, 200, 208, 211, 662, 663, Dipole-dipole interaction, 225, 946
678–680, 934, 942, 952 Dirac, Paul, 22
Deoxygenated red blood cells, 1653 Dirac and Weyl semimetals, 1173, 1175
Deposition, 1607, 1608, 1610 Dirac equation, 188, 201, 202, 227
Diamagnetic susceptibilities, 1596 Dirac Hamiltonian, 201
Diamagnetism, 105, 1596, 1598 Dirac matrices, 201
Dichroic bleaching, 507 Direct coupling, 996–999
Die-upset melt-spun Nd-Fe-B, 1399 Direct magnetoelectric effects, 609
Dielectric targets, 1186 Direct spin-spin interactions, 946
Dielectric tensor, 481, 485, 486 Disordered local moment (DLM), 201,
Differential phase contrast microscopy, 240, 245
1210–1212 Dispersion relation, 193
Index 1685

d 0 magnetism, 914 Earth’s magnetic field, 16, 17


DM interaction, see Dzyaloshinski-Moriya Easy axis, 1374, 1380, 1388, 1404
(DM) interaction material, 1380
DNA molecules, 1658 Easy-plane, 1404
Domain, 1051, 1057, 1206, 1380, 1458 Ecological factors, 1408
analysis, 393–398 Economic development, 1407
branching, 1379 Eddy currents, 665, 1448, 1664
formation, 398–414 brakes, 1420
single, 1060, 1062 classical losses, 1440–1442
structure, 1396 losses, 1413
Domain-size criterion, 525 rotational eddy currents loss, 1448
Domain wall(s), 25, 26 skin effect, 1442–1443
bulk specimens with Q < 1, 419 Edge states, 1060
classical Bloch wall, 417 Edwards-Anderson model, 955
current-driven domain wall motion, Edwards-Anderson order parameter, 954
426–432 ΔE effect, 553–555
dynamics, 420–425 Effective anisotropy constant, 1390
racetrack memory, 1565–1567 Effective magnetic fields, 1312–1314
special walls, 420 Effective magnetoelastic coupling coefficient,
specimens with Q > 1, 419 570, 573
thick films with Q < 1, 418–419 Effective spin hamiltonian, 1304–1306
thin films with Q < 1, 416–418 Efros-Shklovskii Coulomb gap, 935
width, 1394 Eigenvalue problem, 231
3d orbitals, 855 Einstein-de Haas effect, 511
Döring mass, 425 Einstein-de Haas experiment, 17, 21, 47
Double-cone coils, 1667 Elastic energy density, 560
Double exchange interaction, 31, 862 Elastic scattering, 1256
Double-exchange mechanism, 72 Elastic stiffness constants, 560, 578
Double layer, 1602 Elastocaloric effect (ECE), 766, 768
Double perovskites, 879, 884–885 Electric dipole approximation, 481, 485
Droplet theory, 957 Electric field, 436
Drude model, 437 gradient, 123
ferromagnets, 487 Electric field switching in BiFeO3
Dy-free magnets, 1398 bulk single crystals, 605
Dy2 Fe14 B, 84 coupling, 607
Dy2 Ti2 O7 , 97 epitaxial thin films, 606
Dynamical mean field theory (DMFT), 63, magnetization, 602
155, 157, 195, 213, 240, 247, 248 Ni film, 607
Dynamical susceptibility, 237–239 ultrafast THz electric field, 607
Dynamic application with active recoil, 1413, Electric guitar, 1422
1421–1424 Electric motors, 1422
Dynamic application with mechanical recoil, Electric-field control, 173
1412, 1413 Electrical crystal field, 1308
Dyson equation, 195, 199, 222, 225 Electrically detected magnetic resonance
Dzyaloshinskii-Moriya interaction (DMI), 56, (EDMR), 1327, 1328
83, 87–89, 91, 163–165, 189, 405, Electrical/mechanical energy, 1512
419, 862, 865, 872, 891, 903, 944, Electrical steels, 664, 1452
1055, 1056, 1567, 1568 composition, 1452–1453
Dzyaloshinsky-Moriya vector, 222 Fe-Si products, trends in, 1460–1461
non oriented silicon steel sheets, 1455–1457
processing, 1453–1454
E texture, 1454
Early transition metals (ETM), 696 Electrocardiography (ECG), 1639
Earnshaw’s theorem, 1420 Electrochemical cell, 1604
1686 Index

Electrochemistry, 1602–1604, 1606 ODMR, 1328, 1329


Electrode potential, 1603 rotating coordinate system, 1321
Electrodeposition, 1607 time-domain techniques, 1326, 1327
magnetic patterning, 1613 Electron microscopy, 1036, 1207
Electroencephalography (EEG), 1245, 1639, scanning electron microscopy, 1213–1216
1641, 1642, 1648 transmission electron microscopy,
Electromagnetic (EM) 1208–1213
effects, 128 Electron paramagnetic resonance (EPR), 41,
induction, 12 987, 1090, 1123, 1601
induction mechanism, 1639 anisotropic Zeeman interaction,
radiation, 994 1306–1308
smog, 1639 Bloch equations, 1303, 1304
waves, 13 effective spin hamiltonian, 1304–1306
Electromagnetic revolution, 9 FS interaction, 1308, 1309
Charles-Augustin de Coulomb, 9 HF interaction, 1310, 1311
earth’s field, 16, 17 Electron-phonon equilibration time, 506
electricity, 9 Electron-phonon scattering, 437
ferromagnets, 18, 19 Electrophysiological measurements, 1642
magnetostatics, 9, 14 Electrostatic coulomb interactions,
transformation of science, 19 21, 22
Electromagnetism, 11, 18 Electrostatics, 14
Electromagnets, 1087–1088 Elinvar, 553
Helmholtz coils, 1090 Elliott-Yafet spin-flip scattering, 512
resistive and superconducting solenoids, Ellipsoid, 1377
1088–1089 Ellipticity, 483
with magnetic core, 1090–1091 Elongated nanoparticles, 161
Electron Endotheial cells, 1023
beam melting, 1401 Energy
density, 198 barrier model, 1030
discovery, 5 inductive, 1093
holography, 1211–1213 levels, 1305, 1309
intrinsic angular momentum, 21 magnetic, 1092
scattering, 437, 439 product, 1385, 1389, 1392, 1395,
See also Magnetism of electron 1397, 1409
Electron-electron interactions, 59–60 Energy density, 400, 1385
Electronic magnetic moments, 1497 magnetic, 1092
Electronic model systems, 35 Energy-dispersive x-ray spectroscopy, 966
Electronic structure e-p equilibration time, 506
adiabatic dynamics, 204–208 Epitaxial growth, 1159
band structure methods, 196–201 Epitaxy, 1159
calculations, 574 e-p scattering process, 512
dynamical susceptibility, 237–239 Equatorwards, 1636
finite temperature magnetism, 240–248 Equiatomic Fe-Co alloy, 1461–1462
itinerant magnetism of solids, 208–217 Eskolaite, 877
magnon dispersion relations, 231–232 eg states, 577
relativistic effects, 201–204 Eu3+ , 68
spin density functional theory, 190–195 Euler-Lagrange equation, 362
spin spiral calculations, 233–236 Euler method, 377–379
total electronic energy and magnetic Euler’s equation, 416
configuration, 217–230 European Commission, 1021
Electron magnetic resonance European Medicines Agency
EDMR, 1327, 1328 (EMA), 1644
EPR/FMR spectrometers, 1321–1324 Eutectic phase, 1397
frequency domain techniques, 1324, 1325 Excess energy, 1301
Index 1687

Excess loss, 1437, 1438, 1444 F


Bertotti’s model, 1445–1448 Factor of merit, 1450
Pry and Bean model, 1444 Fano resonance, 934
Exchange, 130 Faraday effect, 12, 478, 484, 504, 1480
anisotropy, 1055, 1403 Faraday rotator, 482, 1127–1129
bias, 44, 92, 165, 1049–1051 Fayalite, 897, 898, 902
biquadratic, 64 Fe, 1644, 1654
constant, 1374, 1386 Fe2 B precipitates, 1400
coupling, 189, 221, 241, 242, 927, 1399, α  −Fe16 N2 , 1410
1403, 1404 Fe64 Ni36 , 553
coupling parameters, 220–224 γ − Fe2 O3 , 1644
direct, 56 Fe-Co, 1461–1462, 1464
double, 72 Fe-Cr-Co alloys, 1402
energy, 351–356, 399, 1317, 1318 Fe-Ni, 1463–1467
energy density, 170
permalloy, 1463
field, 1051
permalloy powder cores, 1483
Heisenberg, 1055
hole, 55 powder core, 1483
indirect, 69–72, 1052, 1070 thermal alloys, 1466
interatomic, 54, 56, 60, 65, 74 Fe-rich alloys, 1465
intersublattice, 85 Fe-series anisotropy, 130
intra-atomic, 54, 56, 60, 67–69, 80 Fe-Si, 1453, 1457, 1460
intra-atomic exchange, 106, 125 Fe/Cu (001), 816–820
intrasublattice, 85 Fex Cu1−x /GaAs (001), 836, 837
itinerant, 72–75 Fex Pd1−x /Cu (100), 839
length, 372, 1375 Fex Cu1−x /GaAs (001), 833–836
parameter, 1374 Fex Pd1−x /Cu (100), 837–839
quantum-mechanical origin of, 57–66 Fe100-x Nix , 166
Ruderman-Kittel, 71–72 Fe2 O3 , 165
and spin structure, 83–98 Fe3 O4 , 72, 84, 1644
splitting, 508 Femtomagnetism, 505
stiffness, 56, 83 Ferarris theorem, 1448
Stoner theory, 54, 60, 73, 74, 81 Feripyrophylite, 898
Exchange interactions, 1374 Fermi energy, 208, 225, 243, 630, 949, 952
kinetic exchange, 937 Fermi level, 202, 443, 636, 637, 934, 935,
potential exchange, 937 1214, 1220, 1559
sp-d( f ) exchange interactions, Fermi liquid, 78
938–946 Fermi surface, 441, 630, 634, 638, 639, 642,
Exchange-spring magnets, 1403 643, 647, 652
Excimer lasers, 1181, 1182 Fermi velocity, 438, 630
Exotic tunneling, 456–458 Fermi wavevector, 676
Exposure guidelines, 1667, 1669 Fermi’s Golden rule, 515
Extended Heisenberg Hamiltonian, 244 Fermi-Dirac distribution function, 951
Extended x-ray absorption fine structure Feroxhyte, 868
(EXAFS), 966 Ferric gel, 869
External field energy, 400 Ferrihydite, 869
External magnetic field, 944, 983 Ferrimagnetic, 479
External stress, 369 alloys, 479, 521–523
Extraction magnetometer, 1131–1133 material, 521, 1567
Extraordinary MR (EMR), 1116, 1118 Ferrimagnetism, 24, 84, 1374
Extremely Low Frequency (EFL) compensated, 86
fields, 1666 Ferrite(s), 581, 1086, 1375, 1392–1393
Extrinsic properties, 1384, 1385, 1392, 1398 α−Fe(C), 1390, 1392
Extrusion, 1399, 1412 garnets, 871–872, 1480
printing, 1401 hexagonal ferrites, 873–875
1688 Index

Ferrite(s) (cont.) Fert, Albert, 44


Li, 1479 FFLO state, 629–634, 640, 648, 651, 652
Mg, 1479 Field annealing, 1470
microwave, 1478–1480 Field equivalent noise, 1529
Mn-Zn, 1474–1477 Field free line (FFL), 1650
Ni-Zn, 1477 Field free point (FFP), 1650
Ni-Zn-Cu, 1478 Field gradient force, 1652
orthoferrites, 872–873 Field-cooling, 1051
sintered, 1473–1474 Field/flux sensing, 1529, 1544
spinel, 869–871, 1472–1473 Figure-of-eight coils, 1667, 1668
Ferrite α-Fe(C), 1390, 1392 Filling factor, 1481
Ferritin, 1643, 1645 Film growth methods, 811
Ferroelectric antiferromagnet, Fine structure (FS) interaction, 1308, 1309
601–604, 613 Fine-grained material, 1396
Ferroelectric ferrimagnets, 602, 604 Fine-structure constant, 55
Ferroelectric ferromagnets, 600 Finemet, 1470, 1472
Ferrofluids, 25, 37, 1028, 1029 Finite temperature magnetism
Ferromagnet, 12, 18, 19, 1054, 1596 electronic structure and spin statistics,
superconductor (F/S) systems, 627 244–248
thickness, 1051 longitudinal spin fluctuations, 242–244
Ferromagnetic conductors, 1335 rigid spin approximation, 240–242
Ferromagnetic 3d transition metals, 508 First-principles magnetization-versus-field
Ferromagnetic elements, 28, 44 (FPMVB), 230
Ferromagnetic exchange, 1338 Flat voice coil, 1422
Ferromagnetic films, 572 Flux compression electromagnetic, 1104
Ferromagnetic intermetallic Flux concentrator, 1109
compounds, 33 Flux density, 1376
Ferromagnetic martensites, 599 Flux sensors, 1530
Ferromagnetic oxides, 508 Flux transformers, 1544
Ferromagnetic resonance (FMR), Fluxgate, 1112–1115, 1531, 1532
41, 571, 1026 Foldy-Wouthousen transformation, 201
dipolar interaction, 1312 Food and Drug Administration (FDA), 1021
effective magnetic fields, 1312–1314 Force theorem, 223
exchange energy, 1317, 1318 Form effect, 570
macrospin-like behaviour, 1312 Foucalt imaging, 1210
magnetocrystalline anisotropy, 1312, Four-contact configuration, 465
1316, 1317 Fourier transform holography (FTH), 1236
metals, 1319, 1320 Fourier transformation, 223
microscopic quantum mechanical Fractional quantum-Hall effect (FQHE), 63
approach, 1312 Free electron lasers (FELs), 1231, 1236, 1237
orbital angular momentum, 1312 Free induction decay (FID), 1125
shape anisotropy, 1314, 1315 Freestanding substrates, 1160
spin wave resonance, 1317, 1318 Frequency domain techniques, 1324, 1325
Ferromagnetic semiconductors, 895 Frequency modulation (FM), 1224
Ferromagnetic shape-memory alloys (FSMA), Fresnel coefficients, 489
557–559 Fresnel imaging, 1209
Ferromagnetic superconductors, 648–651 Friedel oscillation, 1052
Ferromagnetic superexchange, 929, 948 Frost’s cycle, 95
Ferromagnetic systems, 523–525 Frozen-magnon calculations, 230
Ferromagnetism, 54, 72, 75, 1374 Frustration, 36, 95, 886, 1070
and magnetotransport, 442–454 geometric, 1061
interatomic exchange and magnetic topologically induced emergent, 1063
order, 55 Functional-magnetic resonance imaging
Ferrous hydroxide, 869 (f-MRI), 1243, 1647, 1648
Index 1689

Functionalised magnetic nanoparticles Ginzburg-Landau expression, 642


(f-MNP), 1663 Ginzburg-Landau parameter, 632, 634
Fused deposition modeling, 1401 Ginzburg-Landau theory, 640
Fused filament fabrication, 1401 Goethite, 868
Future PM materials, 1389 Goodenough, J.B., 861
Frank-van der Merwe growth, 1157, 1183 Goodenough-Kanamori-Anderson rules, 71,
929, 948
Goss texture, 664, 665
G Gradient coils, 1647
GaAs (001) substrates, 811, 812 Gradient force, 1611
Gadolinium, 518 Gradiometer, 1131, 1132, 1543
Gallium Nitride (GaN), 1160, 1161 coils, 1112
Gap Grain alignment, 1400
lumped, 1481 Grain boundary, 1397
spread, 1481–1483 Grain boundary diffusion (GBD), 1398
Garnet ferrites, 30 Grain oriented silicon steels (GOSS), 1457,
Garnets (R3 Fe5 O12 ), 86, 871–872 1462
Gas atomization (GA), 703, 1401 conventional, 1457–1459
Gas phase methods, 1018 cutting, 1459
Gastrulation, 1659 HiB, 1458, 1459
Gauss’s theorem, 265, 359, 360 Grain refinement, 1399
Gd anisotropy, 105 Grain size, 1401
Gd diethylaminetriamine pentaacetate Graphene, 1056, 1170
(Gd DTPA), 1027 Graphene spin logic, 1581, 1582
Gd5 (Six Ge1−x )4 compounds, 767 Grazing incident XRD (GIXRD), 827
General public exposure, 1669 Green function, 199, 226, 227
Generalized Bloch theorem, 233 Green rust, 869
Generalized Fresnel coefficient, 499 Greenalite, 898, 899
Generalized gradient approximation (GGA), Greigite, 894
194, 218, 234, 236, 828 Grossular, 871
Generators, 1391, 1422 Ground manifold, 1305
Geological availability, 1406 Growth modes, 1155–1159
Geomagnetic dynamo, 17 Growth morphology, 1156
Geomagnetic field, 1634–1638 Growth rate, 1165
Geopolitical situation, 1407 Growth temperatures, 813
g-factor Grunberg, Peter, 44
Landè, 1595 Grunerite, 898, 900
spin, 1594 Gutzwiller wave functions, 66, 78
Giant magnetic anisotropy, 171 Gyromagnetic damping, 1479
Giant magnetocaloric effect (GMCE), 1501 Gyromagnetic ratio, 17, 1335
Giant magnetoimpedance (GMI), 464, 1119,
1539, 1540
Giant magnetoresistance (GMR), 39, 44–46, H
446–447, 1059, 1117, 1118, 1344, Halbach array, 1416–1418
1349, 1536, 1555–1557 Halbach cylinder, 1378, 1416, 1417
sensors, 1642 Haldane-Anderson mechanism, 932
Giant orbital paramagnetism, 914 Half-metal, 885
Giant spin Hamiltonian, 988 Half-metallicity, 451
Gibbs free energy, 1155, 1383, 1384 Halides, 890–891
Gilbert damping constant, 382 Hall angle, 458
Gilbert damping constant α, 520 Hall bar, 1115, 1116
Gilbert damping parameter α, 503, 516 Hall conductivity, 205–207
Gilbert relaxation, 1302 Hall constant, 440, 458, 1118
Gilbert, William, 7 Hall current, 438
1690 Index

Hall effect, 188, 438, 440, 458–461, 959, Hemoglobin, 1643, 1646, 1647, 1662
1335, 1563 Hendrika van Leeuwen, 21
sensors, 1534, 1535 Herrings-Kittel formula, 285
Hall probe, 1218 Heteroepitaxial templates, 1160, 1161
Hall resistance measurement, 1568 Heteroepitaxy, 1160
Hall voltage, 458, 465, 1116 Heusler alloys, 88, 213–217, 508, 520,
Hard disk, 1537 559, 1412
Hard ferrite, 1389, 1392–1393 Heusler compounds, 1178, 1506, 1508
Hard steels, 666–667, 1390 Heusler families, 33
Hardness parameter, 1375 Hexagonal ferrites, 30, 873–875, 1392
Hartmann number, 1607 Hexagonal point symmetry, 113
Hartree-Fock approximation, 644 H-field, 13
Hauerite, 893 Hibonite, 873
Hausmannite, 878 High frequency losses, soft
H-coils, 1667, 1668 magnets, 1449
HD-AOS, 505 insulators, 1450–1451
Heat assisted magnetic recording metals, 1450
(HAMR), 1030 High gradient magnetic separation
Heat treatment process, 703, 1394 (HGMS), 1655
Heavy d-shell elements, 689 High magnetic fields, 42
Heavy f-shell elements, 690 High-field magnet facilities
Heavy fermions, 83 DC-magnets and facilities, 1094–1099
Heavy p-shell elements, 689 Megagauss magnetic fields, 1102–1104
Heavy-fermion compounds performance limitations and magnet
basic properties, 776 classification, 1091–1095
electron mass, 771 pulsed magnets and facilities, 1099–1102
energy scale, 773 High-field resistive magnets, 1095
fermi level, 772 High-gradient magnetic separators (HGMS),
fermi-liquid, 771 1616, 1618
groups, 778 High-harmonic generation (HHG), 496
Kondo effect, 772 High-resolution transmission electron
Kondo-lattice compounds, 773 microscope (HRTEM), 816
magnetic interactions, 778 High-spin low-spin transitions, 69
microscopic level, 772 High-temperature superconductivity (HTSC),
non-superconducting, 775 94, 95
Quantum phase transition, 773 High-temperature superconductors (HTC),
Sommerfeld coefficient, 771 1175, 1176
unconventional superconductivity, 778 Highly-oriented pyrolytic graphite (HOPG),
uranium/cerium intermetallics, 774 689, 1170
vegetable, 774 History of magnetism, 4–10
Hedenbergite, 898, 901 compass, 5–7
Heisenberg exchange, 1055, 1335 Lucretius, 5
interaction, 1265 magnetic attraction to Thales, 5
Heisenberg hamiltonian, 22, 220, 221, 862 modern science, 7
Heisenberg model, 34, 63–65, 92, 97, 105, Holding components, 1419
165, 349 Holding magnet, 1419
Heisenberg picture, 520 Homoepitaxy, 1158
Heisenberg, Werner, 22 Hopping (interatomic), 56, 58, 72–74, 82,
Heitler-London approximation, 56 93, 95
Helicity-dependent multiple-pulse Horseshoe magnets, 1391
mechanism, 524 Horseshoes, 4, 8, 15
Helimagnetism, 87 Hot deformation, 1388, 1399, 1400
Helmholtz coils, 1090 Hot extrusion, 1412
Hematite (α-Fe2 O3 ), 87, 864–865, 1374 Hot isostatic pressing (HIP), 703
Index 1691

Hot pressing (HP), 703 Interfacial Dzyaloshinskii-Moriya


Hubbard band, 861, 876 interaction, 526
Hubbard model, 35, 65–66, 647 Interference electron microscopy, 1034
Hubbards bands, 934 Interlayer coupling, 283
Human exposure, 1666 Interlayer exchange coupling, 1052, 1054
Hund’s rules, 22, 67, 68, 130, 451, 660, 858 Intermetallic compounds, 3d-4f type
Hybrid magnets, 1098 anisotropy parameters, 744, 749
Hybridization gap, 935 antiferromagnetic behavior, 743
Hybridization-type ligand fields, 150 antiferromagnetic coupling, 754
Hydrogen, 754 antiferromagnetism, 753
atoms, 1026 binary and ternary systems, 727
liquefaction, 1496 Co-Co exchange interactions, 731
Hydrogen disproportionation desorption collinear antiferromagnetic
recombination (HDDR) process, structures, 752
765, 1387 cryogenic temperatures, 749
Hydrostatic pressure, 766 crystal electric field parameters, 750
Hydrothermal synthesis, 1020 crystal structure, 752
Hyperfine (HF) interaction, 1310, 1311 hard magnetic and magnetocaloric
Hyperthermia, 1664 applications, 727
Hysteresis, 18, 43, 1382, 1384 hydrogen insertion, 754
Hysteresis curves, 394, 396, 1383 magnetic moments, 743
Hysteresis loop, 18, 19, 25, 27, 396, 990, magnetic phase diagram, 749, 750
1024, 1382 magnetic structure, 751
Hysteresis loss, 1436 magnetocaloric effect, 737
rotational hysteresis loss, 1448 magnetocrystalline anisotropy, 743, 749
Mössbauer spectroscopy analysis, 744
Nd2 Fe14 B, 749
I Nd5 Fe17 and Sm5 Fe17 , 727
Ilmenite, 877 Nd6 Fe13 Si compound, 753
Ilvaite, 898, 901 neutron diffraction experiments, 731,
Immunoassays, 1656 744, 753
Impurity band, 914 Néel temperature, 754
Impurity complexes, 966–967 planar anisotropy, 743
Independent-electron approximation, 61 R2 Co14 B type compounds, 747
Indirect coupling, 999–1001 R2 Co17 compounds, 741
Indirect exchange, 1052 R2 Co7 compounds, 728
Induced anisotropy, 400 RFe5 , 731
Induction, 1111, 1113, 1131–1136, 1376 R2 Fe14 B compounds, 745
energy, 1093 R2 Fe14 C type compounds, 748
sensors, 1530, 1531 R2 Fe17 compounds, 739
Inelastic neutron scattering technique, R2 Ni7 compounds, 729
1265–1266 R5 Fe17 compounds, 730
Ingot materials, 1396 R6 Co13-x Ax compounds, 756
Initial magnetization curve, 1382 R6 Fe13-x Ax compounds, 753, 755
Injection molding, 1400, 1401 R6 M13-x Ax Hy (M=Fe, Co) compounds,
In-plane anisotropy, 112 756
In-situ monitoring R2 Fe14 B compounds, 749
thin film growth, 1168, 1169 R2 Fe17 compounds, 743
Interaction domains, 1381, 1387, 1396 R2 Fe17 ferromagnetic compounds,
Interaction parameter, 1607 743
Interatomic exchange, 54, 56, 74 R2 M14 B compounds, 744
Interface exchange bias, 504 R2 M17 compounds, 737
Interface spin current, 1338 R6 M13 A compounds, 754
Interface states, 1057 R6 M13 A phases, 750
1692 Index

Intermetallic compounds, 3d-4f type (cont.) Iron hydroxides


RCo5 compounds, 732, 736, 737 ferrihydrite and ferric gel, 868–869
RCo5 compounds, 734, 737 ferrous hydroxide, 869
RM5 compounds, 731, 734‘RNi5 goethite and oxyhydroxides, 868
compounds, 735–737 Iron oxide(s), 863, 867–868, 1020
RNi5 compounds, 734 hematite, 864–865
spin reorientation, 743, 750 iron hydroxides, 868–869
superstructure reflexions, 743 maghemite, 867
temperature dependence, 751 magnetite, 866–867
transport properties, 754 Iron oxide nanoparticle (IONP), 1644–1646
YCo5 and isotype compounds, 731 Ising model, 34, 92
International Agency for Research on Cancer Island growth, 1158
(IARC), 1666 Isothermal entropy, 1507
International Commission on Non-Ionizing Isotropic exchange, 1317
Radiation Protection (ICNIRP), Isotropic ferromagnets, 1283
1667, 1669 Isotropic magnetostriction, 564–565
Interstitial atoms, 1404, 1411 Isotropic nuclear Zeeman effect, 1310
Interstitially modified, 1404 Isotropic stress, 568
Interstitial Mn ions, 964–965 Itinerant anisotropy, 151–154
Intersublattice exchange, 85 Itinerant electron metamagnetism (IEM), 1505
Intra-atomic exchange, 54, 56, 67–69, 858 Itinerant exchange, 72–75
Intrasublattice exchange, 85 Itinerant magnetism, 130
Intrinsic crystal fields, 139 Itinerant magnetism of solids, 208
Intrinsic Loss Power (ILP), 1025 Heusler alloys, 213–217
Intrinsic magnetic lengths, 1375 Slater-Pauling curve, 211–213
Intrinsic magnetic properties, 1384, 1392, 1394 Stoner model, 208–211
Intrinsic spin wave, 1285 Itinerant magnets, 117, 145, 150, 576
Invar effects, 553
Invar Fe64 Ni36 , 27
J
Inverse Faraday effect (IFE), 484, 504,
Jahn-Teller effect, 559, 856, 857, 931, 932
523, 525
Jellium model, 72
Inverse magnetoelectric effect, 609–613
Johnson noise, 1108
Inverse magnetostrictive effect, 567
Jones matrix, 491, 493
Inverse spin Hall effect, 462
formalism, 482
Inverse square laws, 9
Josephson junction (JJ), 1119, 1120, 1532
Ion beam deposition (IBD)
Joule, James, 18
geometry, 1193, 1194
Joule heating, 1092
source, 1193
Joule magnetostriction, 550, 551, 557
source operation, 1195
target, 1194, 1195
Ion beam sputtering K
applications, 1195, 1196 Kanamori, J., 861
atom kinetic energy, 1192 Karelianite, 876, 877
IBD geometry, 1193, 1194 Karplus-Luttinger mechanism, 205
IBD source, 1193 Kasuya’s model, 946
IBD source operation, 1195 Kelvin force, 37, 1610, 1611
IBD target, 1194, 1195 Kerr ellipticity, 492
Ionic radii, 853 Kerr microscopy, 498, 1207, 1381
Iron (Fe), 132, 148, 1452, 1642, 1643 Kerr rotation, 483
carbonyl, 1482 Kerr vector, 492, 493
loss, 1437 Kinetic(s)
powder cores, 1482–1483 Butler-Volmer, 1602
Iron-based superconductors, 627 exchange, 937
Iron-cobalt-based nanostructures, 698 processes, 1155
Index 1693

Kirkendall effect, 1015 Laser ablation, 1182


Kitaev magnetism, 888 Laser-induced demagnetization, 479
Kittel equation, 286 conservation of angular momentum,
Knudsen cell, 1167, 1171 510–512
Kohler’s rule, 441, 442 experimental demonstration, 505–508
Kohn anomalies, 676 observations in laser-induced fs
Kohn-Sham equation, 154, 192, 196, 197 demagnetization, 508–510
Kohn-Sham orbitals, 194 quantitative understanding, 517–521
Kondo coupling, 936 theories for femtosecond demagnetization,
Kondo effect, 34, 81, 94, 772, 926, 928, 934, 512–517
935, 937, 1067–1069 Laser-induced growth of ferromagnetic
Kondo insulators, 935 order, 504
Kondo temperature, 934 Laser-induced magnetization dynamics,
Korringa-Kohn-Rostoker (KKR), 199 478, 502
Kristallfeld, 1374 all-optical switching of magnetization,
Kronecker’s delta, 1605 521–527
Kittel formula, 1314 laser-pulse-excited spin currents, 527–534
ultrafast laser-induced loss of magnetic
order, 505–521
L Laser-induced spin currents, 505
L10 -FeCo, 1411 Laser-pulse-excited spin currents, 527
L10 -FeNi, 1410 optical spin-transfer torque, 530–532
L10 -MnAl, 1411 optical THz spin wave excitation,
L10 FePt and CoPt, 1404 532–534
La(Fe1-x Six )13 alloys, 767 optically-induced spin transfer, 528–530
Lab-on-chip (LOC) Lattice-matched epitaxial growth
magnetic mixing, 1657 strategies, 1163
magnetic transportation, 1657 Lattice mismatch, 1161, 1162
magnetic trapping and separation, 1656 Lattice strain, 585
Ladder operators, see Spin-flip operators Layer-by-layer growth, 1156, 1157
Lagrange multipliers, 378 Layer-specific measurements, 493
Laihunite, 898 Layer-specific MOKE, 492–495, 502
LaMnO3 , 72 Layered magnetic materials, 1169
Landé factor, 350, 940, 943 LDA+DMFT, 189
Landè g-factor, 1595 LDA+U, 195
Landau level, 930 Lehmann spectral representation, 198
Landau theory, 596 Length, magnetic, 1105
Landau-Lifshitz Bloch (LLB) equations, 515 Lenz’s law, 1596
Landau-Lifshitz-Gilbert (LLG) equation, Lepidocrocite, 868
502, 504, 1015, 1026, 1312, Levitation, 1420, 1653
1334–1337, 1347 Liebenbergite, 902
Landau-Lifshitz-Gilbert-Slonczewski (LLGS) Ligand field, 855, 1070
equation, 422, 426, 431 interaction, 27
Landau-type potential, 1383 Light p-shell elements, 689
Landau-Zener-Stückelberg (LZS) model, Linear magnetoresistance, 440
989–991 Linear response theory, 210
Landau–Zener tunnel probability, 990 Linear sensors, 1536
Lande factor, 1307 Linearized augmented plane-wave
Langevin function fit, 1014 (LAPW), 197, 833
Lanthanide (III) ions, 984 Linneite, 894
Laplace’s equation, 672 Liposomes, 1659
Large scale coherence, 274 Liquid phase syntheses, 1018, 1019
Large scale entanglement, 274–276 Liquid-phase sintering, 1397
Larmor frequency, 1123–1126, 1239, 1300 Lithium ferrite, 870
1694 Index

Lithographically-patterned 2D ferromagnetic Rayleigh loops, 1439–1440


arrays rotational losses, 1448–1449
artificial spin-ice structures, 1061, 1063 Steimetz model, 1438
nano/microdisc arrays, hysteresis of, 1060 Low-noise amplifier (LNA), 1327
LMTO, 197 Low Si alloys, 1460
Load line, 1412 Low-spin state, 858
Local dissipation of angular momentum, 529 LSDA+DMFT, 236
Local spin-density approximation (LSDA), LSDA+U, 63, 78–80, 236
193–195, 202, 211, 218, 223, Luttinger liquid, 93, 94, 97
234, 236 Luttinger spin-orbit interaction, 929
Local-density approximation to
density-functional theory
(LSDA DFT), 61 M
Localization, 62, 65, 71, 72, 74, 82, 83, 86, 90 Macroscopic, 392
Localized-electron magnetism, 29, 31 equation of motion, 1301, 1302
Lodestone, 4–7, 16, 24, 1374, 1387 method, 1312
Logarithmic time dependence, 1382 micromagnetic free energy, 1302
Longitudinal configuration, 485 quantity, 1541
Longitudinal magnetization dynamics, 516 tunneling, 271–274
relaxation, 1647 Macrospin(s), 1013, 1034, 1341
time, 994 model, 1013, 1034
Longitudinal spin fluctuations, 242–244 Maghemite, 867, 870
Lorentz, Hendrik, 13 Magic mangle, 1417
Lorentz force, 438, 458, 1116, 1129, 1422, Magnetic 3d transition metals
1605, 1607, 1611–1613, 1665 Co/GaAs (001), 819–824
density, 1092 Cox Mn1−x /GaAs (001), 829–832
Lorentz microscopy, 1208–1209 Cr/GaAs (001), 812, 813
differential phase contrast microscopy, Fe/Cu (001), 816–820
1210–1212 Fex Cu1−x /GaAs (001), 836, 837
Foucalt imaging, 1210 Fex Pd1−x /Cu (100), 839
Fresnel imaging, 1209 Fex Cu1−x /GaAs (001), 833–836
Loss Fex Pd1−x /Cu (100), 837–839
anomalous, 1437 Mn/GaAs (001), 814–816
classical, 1440–1442 Ni/GaAs (001), 824–826
eddy current, 1440 Py/GaAs (001), 826–829
excess, 1437, 1444–1448 Magnetic actuation, 1653
hysteresis, 1436, 1438 biology and biomedical applications,
Preisach model, 1438 1652, 1653
rotational, 1448–1449 field gradient force, 1652
separation, 1436 levitation, 1653
skin effect, 1442–1444 lab-on-a-chip, 1655–1657
units, 1437 magnetic orientation, 1662, 1663
Loudspeakers, 1421 magnetic trapping, 1654–1655
Low carbon steel, 1452 magnetic tweezers, 1657–1661
Low cobalt alloys, 1463 magnetoporation, 1661
Low-energy electron diffraction (LEED), 812, superconducting coils, 1653
818, 1168 torque, 1652
Lowering operators, see Spin-flip operators Magnetic after effect, 1381
Lowest-order anisotropies, 108–109 Magnetically activated cell sorting
Lowest unoccupied molecular orbital (MACS), 1654
(LUMO), 1222 Magnetically compensated hinges, 1420
Low frequency losses, soft magnets Magnetically induced cell destruction,
eddy currents, 1440–1444 1660, 1661
excess loss, 1444–1448 Magnetically-ordered compounds, 34
Index 1695

Magnetically stabilised bed (MSB) Magnetic-field-induced strains


reactors, 1621 (MFIS), 705
Magnetic and kinetic moments, 1299 Magnetic fixation, 1650
Magnetic anisometer, 1137 Magnetic fluids, 1028
Magnetic anisotropy, 104, 1070, 1137, 1314, Magnetic flux, 1131
1374, 1394 Magnetic force(s)
Magnetic attraction, 5 Kelvin, 1605, 1610
Magnetic band theory, 23 field-gradient, 1605, 1610, 1612, 1616
Magnetic beads, 1644, 1654, 1656–1659 Helmholtz, 1611
Magnetic bearings, 1420 Kelvin, 1612
Magnetic bilayer spin-valve stacks, 46 Lorentz, 1605, 1607, 1611, 1613
Magnetic carriers, 1618 Maxwell stress, 1605
Magnetic cell separation, 1654
in non-uniform magnetic field,
Magnetic charge, 9, 14
1610–1615
Magnetic circuit, 1091, 1391
Magnetic circular dichroism (MCD), 116, theorem, 225
481, 522 in uniform magnetic field, 1605–1610
Magnetic constant, 9 Magnetic-force meter, 1129
Magnetic contrast, 482 Magnetic force microscopy (MFM), 689, 1036,
Magnetic cooling engines, 1509–1511 1140, 1217, 1223–1226, 1396
Magnetic core, 1090–1091 Magnetic free energy, 1312
Magnetic couplings, 1420 Magnetic frustration, 95
Magnetic diffraction, 1269 Magnetic gears, 1420
with neutrons, 1268–1276 Magnetic Gibbs free energy
with x-rays, 1276–1280 crystal anisotropy energy, 363–366
Magnetic dipole, 10 exchange energy, 351–356
transitions, 1302, 1306 external stress, 369
Magnetic domain(s) magnetoelastic coupling energy, 367–369
domain classification, 409–414 magnetostatic energy, 357–363
domain walls, 414–432, 1576 magnetostatics, 356–357
domains and domain analysis, 393–398 magnetostrictive self energy, 369–371
magnetic energies, 399–403 spin, magnetic moment and magnetization,
supplementary domains, 406 350–351
Magnetic drug delivery, 1023 spontaneous magnetostrictive
Magnetic electrodes, 1562 deformation, 367
Magnetic energy(ies), 399, 1092 Zeeman energy, 357
anisotropy energy, 400 Magnetic hard drive memory
density, 1092 analog data storage devices, 1554
domain wall energy, 403 giant magnetoresistance, 1556, 1557
exchange energy, 399 RKKY coupling, 1556, 1557
external field energy, 400 road map, 1555, 1556
magneto-elastic interaction energy, 402 tunneling magnetoresistance, 1557, 1559
magnetostrictive self energy, 401–402 Magnetic heating, 1664–1665
stray field energy, 401 Magnetic hemofiltration, 1654, 1655
Magnetic equator, 1634 Magnetic Heusler compounds
Magnetic excitations, 1286 atomic interactions, 722
Magnetic field(s), 9, 11, 1556, 1594–1596 atoms, 717
effects, 1105–1107, 1599–1601 band gap, 717
generation, 1084–1104 band Jahn–Teller effect, 724, 725
inversion, 436 borderline materials, 717
measurements, 1109–1130 chemical notation, 717
Magnetic field generation, 1084 chemical synthesis, 723
electromagnets, 1087–1091 crystal structure, 724
high-field magnet facilities, 1091–1104 cubic compounds, 725
permanent magnets, 1085–1087 electron compounds, 717
1696 Index

Magnetic Heusler compounds (cont.) Magnetic measurements, 1104


electronic band structure, 725 bulk magnetic measurements, 1130–1139
electronic states, 722 calibration and metrology, 1141–1143
electronic structure, 717 fabrication and cost, 1107
Fermi energy, 722, 725, 726 flux concentration, modulation, null-
ferrimagnetic order, 722 detection, 1109
half-Heusler compounds, 717 magnetic field effects, 1105–1107
magnetic shape memory compounds, 725 magnetic field measurements, 1109–1130
manganese containing materials, 717 noise, 1108–1109
materials class, 717 signal properties vs. sensing
mechanism, 717
requirements, 1107
non-centrosymmetric structure, 727
stray field mapping, 1139–1141
phase transition, 727
pseudo-gap, 725 technical environment, 1107
semiconducting rare-earth, 717 technical parameters, 1107
spin reorientation, 727 Magnetic metallic films (MMFs), 810
structural parameters, 725 Magnetic microstructure analysis, 392
tetragonal compounds, 725 Magnetic mixing, 1657
tetragonal distortion, 724, 725 Magnetic moment, 10, 116, 130, 350, 1595
valence electrons, 722, 726 density, 669
Wyckoff position, 722 per atom, 1033
zinc-blende crystal structure, 721 Magnetic monopoles, 97
Magnetic history, 1374, 1381 Magnetic multilayer(s), 1049
Magnetic hyperthermia, 1024–1026 heterostructures, 1556
Magnetic imaging MPI, 1649–1651 Magnetic nanoparticles (NPs)
Magnetic levitation, 1611 biomedical applications, 1644–1646
Magnetic linear dichroism (MLD), 496 formation, 1017, 1018
Magnetic logic Magnetic nanoparticles (NPs) anisotropy, 1036
electronics rely, 1573 antiphase boundaries, 1016
magnetic domain wall, 1576 aqueous syntheses, 1020
magnetic tunnel junction logic biaxial and cubic anisotropy, 1013
devices, 1578 biomedical applications, 1012, 1020, 1021
nanomagnet logic, 1574–1576 bulk material, 1015
nonvolatile memory elements, 1573 bulk parameters, 1013
notations/symbols, 1574 characteristics, 1021, 1022
Magnetic mangle, 1417 chemical composition, 1016
Magnetic materials Alnico magnets, 28 dipolar interactions, 1017
amorphous magnets, 35–36 electron biprism, 1036
computational methods, 43 electron microscopy, 1036
crystal structures, 28 Fe and Co oxides, 1015
energy product, 28 Fe3 O4 , 1016
experimental methods, 41–43 ferrofluids, 1028, 1029
intermetallic compounds, 32, 33 ferromagnetic material, 1012
invar Fe64 Ni36 , 27 flux, 1036
magnetic fine particles, 36, 37 FMR measurements, 1035
magnetic oxides, 29–31 hyperthermia applications, 1017,
magnetic recording, 38, 39 1024–1026
materials preparation, 40–41 iron oxide, 1016
model systems, 34–35 Kirkendall effect, 1015
multielectron atoms and ions, 28 Langevin function fit, 1014
permalloy Fe20 Ni80 , 27 Larmor precession frequency, 1014
single crystals/thin films, 40 liquid phase syntheses, 1018, 1019
thin films, 41 long-range magnetostatic
traditional magnetic materials, 27 interactions, 1012
transformation, 40 macrospins, 1012, 1013, 1015
Index 1697

maghemite, 1027 perovskites, 878–885


magnetic fluids, 1012 pyrochlores, 885–886
magnetic hyperthermia, 1014 Magnetic particle hyperthermia, 1664, 1665
magnetic moment, 1012 Magnetic particle imaging (MPI), 1026,
magnetic recording media, 1029, 1030 1649–1651
magnetism, 1035 Magnetic phase transition, 504
magnetite, 1027 Magnetic point groups, 112
magnetoresistance, 1035, 1036 Magnetic pressure, 1091
manipulation, 1023, 1024 Magnetic principal axis, 988
metal oxide, 1015 Magnetic random-access memory (MRAM),
MFM, 1036 1335, 1537
microSQUIDs, 1035 ferromagnetic electrodes, 1560
monodomain, 1012, 1013 memory cells, 1560
MPI, 1026 silicon-based RAMs, 1559
MRI, 1026–1028 SOT-MRAMs, 1562–1564
nanocrystalline materials, 1017 STT-MRAM, 1560, 1562
negative surface anisotropy favors, 1015 Magnetic recording, 38, 39, 105
non-aqueous syntheses, 1018, 1019 media, 1029, 1030
non-magnetic factors, 1036 Magnetic resonance force microscopy
Néel-Brown model, 1035, 1036 (MRFM), 1218
physics-oriented efforts, 1037 Magnetic resonance imaging (MRI), 1022,
positive surface anisotropy favors, 1015 1023, 1026–1028, 1089, 1238,
quantum mechanical exchange, 1017 1243, 1415
separation, 1023, 1024 cardiovascular imaging, 1647
shell, 1015 f-MRI, 1647, 1648
superferromagnets, 1017 gradient coils, 1647
superparamagnetic particles, 1014 longitudinal relaxation, 1647
superparamagnets, 1015 musculoskeletal imaging, 1647
surface anisotropy, 1015 NMR, 1646
surface magnetization, 1016 real-time MRI, 1647
surface spins, 1013 RF magnetic field, 1646
surfactant/polymer coating, 1015 static magnetic field sources, 1648, 1649
Magnetic navigation 1634 transverse relaxation, 1647
geomagnetic field, 1634 Magnetic resonance navigation
magnetotatic bacteria, 1635, 1636 (MRN), 1655
See also Avian magnetic navigation Magnetic resonance of electrons
Magnetic neutron diffraction, 1257–1260 EPR, 1302
polarized neutron reflectometry, 1261 exchange and dipolar interactions, 1302
polarized neutron techniques, 1260–1261 FMR, 1302
Magnetic object, 1446 local environment, 1302
Magnetic order parameter, 34 macroscopic equation of motion,
Magnetic ordering temperature, 862 1301, 1302
Magnetic orientation, 1662, 1663 microscopic equation of motion,
Magnetic oxides 1299–1301
monolayers, 909–914 Magnetic scale, 1138, 1139
3d oxides, 876–878 Magnetic scattering
3d oxides, 876 inelastic neutron scattering technique,
4d and 5d oxides, 887 1265–1266
4f oxides, 888 magnetic diffraction, with neutrons,
4d and 5d oxides, 887, 888 1268–1276
4f oxides, 888, 889 magnetic diffraction, with x-rays,
5f oxides and related compounds, 889 1276–1280
5f oxides and related compounds, magnetic neutron diffraction technique,
889, 890 1257–1261
1698 Index

Magnetic scattering (cont.) Magnetic voice recording, 19


resonant inelastic x-ray scattering Magnetic water treatment, 1614
technique, 1267–1268 Magnetism, 660, 1070
resonant magnetic x-ray diffraction and electricity, 10
technique, 1262–1265 charged particles, 5
spin dynamics, 1282–1289 chromium, 668–669
Magnetic semiconductors, 508 experimental science, 40
Magnetic sensors, 1420 ferromagnetism, 10
characteristics, 1529 iron, cobalt and nickel, 662–664
electrical current, 1528 iron, steels and iron-based alloys, 664–667
electrical signal, 1528 manganese, 667
field/flux sensing, 1529 1930 Solvay Conference, 24
nervous system, 1528 spin density waves, 669, 670
parameters, 1528, 1529 See also History of magnetismhistory
physical quantity, 1528 Magnetism and superconductivity, 626
Magnetic separation, 1022, 1616–1618, 1656 antiferromagnetic order, 640–648
biological, 1621 ferromagnetic superconductors, 648–651
(HGMS)y, 1618 Ising superconductors, 638–639
high gradient, 1618 paramagnetic limit and non-uniform FFLO
magnetic carriers, 1621 superconducting state, 634
magnetic nanoparticles, 1618 Zeeman field and spin-orbit interaction,
MSB reactors, 1621 635–638
recycling, 1618 Magnetism of electron, 20
waste-water, 1619 angular momentum, 21
Magnetic shape memory (MSM), 705, classical physics, 20
716–718, 720 discovery, 20
Magnetic steels, 1454 Magnetite, 24, 866–867, 1387, 1473
Magnetic substances, 110 Magnetite-based magnetoreception, 1637
Magnetic susceptibility, 1595–1597 Magnetization, 10, 11, 1374, 1375, 1595
of liquids, 1598 angles, 106
Magnetic switching curve, 392, 393, 1382
orthogonal spin-torque driven, 1342–1344 precession, 504
spin-torque driven anti-damping, process, 394, 397, 1384
1339–1342 vector, 393
Magnetic tape recording, 38 Magnetization dynamics, 382
Magnetic targeted drug delivery, 1654, 1655 spin transfer torque, magnetic switching
Magnetic thin-film technology, 46, 167 and oscillations, 1337–1344
Magnetic torque, 633, 634 spin-torque oscillators, 1344–1347
Magnetic tracer, 1649 synchrotron and femtosecond-laser based
Magnetic transition metals (MTM), 696 time-resolved spin dynamics,
Magnetic transportation, 1657 1347–1351
Magnetic trapping (non-lab-on-chip), 1656 ultrafast spin transfer torques, 1351–1353
magnetic cell separation, 1654 Magnetization-induced second-harmonic
magnetic hemofiltration, 1654, 1655 generation (MSHG), 496, 505,
magnetic targeted drug delivery, 1654, 506, 529
1655 Magnetizer, 1388
Magnetic tunnel junction (MTJ), 45, 1065, Magneto-active substrates, 1659
1117, 1339, 1341, 1565 Magneto-aerotactic mechanisms, 1636
logic devices, 1578 Magneto-biology
Magnetic tweezers exposure guidelines, 1667, 1669
cell destruction, 1660, 1661 magnetic fixation, 1650
mechanical stimulation, 1659, 1660 magnetic heating, 1664
physical properties, 1657–1659 magnetic imaging, 1646–1651
Magnetic viscosity, 1381, 1382, 1387 magnetic nanoparticles, 1644–1646
Index 1699

TMS, 1667, 1668 Magnetochemistry, 1599


See also Magnetic actuation radicals, 1599
Magneto-Coulomb effect, 1068 single-triplet interconversion, 1599
Magneto-crystalline anisotropy (MCA), spin states, 1599
224–230 Magnetocrystalline anisotropy (MCA), 26,
Magneto-electric sensors, 1540, 1541 105–107, 159, 160, 355, 400, 402,
Magneto-optical (MO) Kerr effect (MOKE), 574, 1059, 1068, 1316, 1317
42, 43, 483, 484, 497, 529, 818 Magnetoelastic and elasticity data, 577–578
different configurations, 489–492 Magnetoelastic anisotropy, 166, 1316
layer-specific MOKE, 492–495 Magnetoelastic coupling, 556
MO spectroscopy, 495–496 coefficients, 552, 560
MOKE microscopy, 496–497 energy, 367–369
Magneto-optic Kerr microscopy, 42 in films, 571–574
Magneto-optical parameter, 482 Magnetoelastic effects, 673–675
Magneto-optics anisotropic magnetostriction (Joule
macroscopic perspective, 480–485 magnetostriction), 551
microscopic understanding, 485–488 ΔE effect, 553–555
MOKE, 488–497 FSMA, 557–559
ultrafast magnetization dynamics, 497–502 in films, 569, 570
Magnetocaloric(s), 105 magnetomechanical damping, 555, 556
cooling, 1400 magnetovolume effects, spontaneous
intermetallics, 766–771 magnetostriction, forced
Magnetocaloric effect (MCE), 766, 768 magnetostriction and invar
adiabatic temperature, 1491 effects, 552
ADR materials, 1495, 1496 Matteucci effect, 557
anisotropy constants, 1491 spin-state transition, in
barocaloric effect, 1492 La(1−x) Srx CoO3−δ (x ≥ 0.3),
caloric material properties, 1494 559
conventional, 1492 superconductors, magnetostriction in, 559
cooling cycle, 1493 Villari effect, 553
efficiency, 1494 Wiedemann effect, 556, 557
entropies, 1490 Magnetoelastic energy density, 560
equilibrium thermodynamics, 1490 Magnetoelastic interaction energy, 402
ferromagnetic state, 1494 Magnetoelasticity, 571
free energy, 1490, 1491 and Joule magnetostriction, 560–567
gas-based cooling cycle, 1493 Magnetoelectric(s)
direct magnetoelectric voltage
gas-based cooling devices, 1492
coefficient, 610
gas-free cooling engine, 1494
laminates, 1129, 1130
isothermal entropy, 1491 magnetoelectric coefficients, 597
magnetic cooling, 1494, 1495 piezoelectric effect, 596
magnetic field, 1490 sensors, 609
magnetic phase transitions, 1495 solid-state materials, 596
magnetic refrigerants, 1494 spin-orbit logic, 1583, 1584
Maxwell relation, 1490 Magnetoelectric effects
moments, 1490 direct magnetoelectric effect, 609
NDR, 1497, 1498 in composite magntoelectrics, 609
phase transition, 1493 in ferroelectric materials, 597
in time-asymmetric media, 597
phononic subsystem, 1493 inverse magnetoelectric effect, 609–613
temperature, 1490 Landau theory, 596
temperature change, 1494 magnetically ordered single-phase
thermodynamics, 1490 materials, 597
transition materials, 1493 magnetoelectric and multiferroic
volatility, 1494 materials, 608
See also Room temperature at room temperature, 608
Magnetocardiography (MCG), 1639 single-phase materials, 597
1700 Index

Magnetoelectrochemistry, 1601–1615 Magnetostriction data


Magnetoencephalography (MEG), 1244–1245, amorphous Fe alloys, 579
1639, 1641, 1642, 1648 Bi, 582
Magnetofossils, 1636 Fe-Ga (Galfenol), Fe-Ge, FeAl, Fe-Si,
Magnetohydrodynamics (MHD), 1594, 1601, Fe-Ga-Al, Fe-Ga-Ge alloys,
1603, 1606–1608 579–580
characteristic MHD numbers, 1606 oxide materials, 585
micro-MHD, 1607, 1608 paramagnetic metals and alloys, 582
Magnetoimpedance (MI) giant, 464, 1119 Tb, Dy and Ho, 582
Magnetomechanical damping, 555, 556 TbFe2 (Terfenol) and Tb27 Dy73 Fe2
Magnetometer, 1542 (Terfenol-D), 584
cantilever, 1137–1138 Magnetostrictive alloys, 551
extraction, 1131 Magnetostrictive interaction, 408
fluxgate, 1113 Magnetostrictive self energy, 369–371,
induction, 1131–1136 401–402
pulsed-field, 1134–1135 Magnetostrictive strain tensor
torque, 1136–1139 cubic case, 560–563
vibrating sample, 1133–1134 hexagonal case, 565–566
Magnetophoresis, 1614, 1618, 1623, Magnetotatic bacteria, 1635, 1636
1656, 1657 Magnetotaxis, 1635, 1636
Magnetophosphenes, 1666 Magnetotransport (MTR), 43
Magnetoplumbite, 873 anisotropic magnetoresistance, planar
Magnetoporation, 1661 Hall effect and two-current model,
Magnetoreception, 1600 442, 445
in birds, 1636 CMR, 448–451
magnetite-based, 1637 exotic tunneling, 456–458
radical-pair based, 1637–1639 GMR, 446–447
Magnetoreceptors, 1601 Hall effect, 458–460
Magnetoresistance (MR), 439, 883, 1023, magnetoimpedance, 464
1035, 1057, 1116, 1118 measurements, 464–466
anisotropic, 1116, 1118 OMAR, 454
giant, 1059, 1116–1118 PMR, 454
in semiconductors and semimetals, quantum transport, 454–456
440–442 spin currents, 460–462
tunnel, 1065, 1067, 1117, 1118 TMR, 451–454
Magnetoresistive sensors, 1544 Magnetrode, 1642
Magnetoresitive conductor, 1327 Magnetron, 1418
Magnetosomes, 1635 operation, 1187, 1188
Magnetostatic(s), 9, 14, 356–357 plasma, 1187
anisotropy, 159, 161 sources, 1187, 1188
boundary value problem, 361–362 Magnetron sputtering, 811
waves, 1318 alloy films, 1191, 1192
Magnetostatic energy, 27, 357, 360–363, 375, chimneys, 1189, 1190
1313, 1385, 1611 gas molecules, 1186
demagnetizing field, 358–359 magnetron operation, 1187, 1188
magnetic scalar potential, 359 magnetron sources, 1187, 1188
magnetostatic boundary value problem, material targets, 1186
361–362 planar magnetron, 1186
Magnetostriction, 18, 33, 166, 570, 1462 reactive, 1190, 1191
of polycrystalline cubic materials, 564–565 RF, 1190, 1191
of polycrystalline hexagonal materials, 566 shutters, 1189, 1190
and stress-induced magnetic anisotropy, Magnets, 1650
567–569 Bitter-type, 1097
in superconductors, 559 DC, 1094–1099
Index 1701

electro, 1087–1091 Medical magnetic imaging, 1238


high-field magnet facilities, 1091–1104 magnetic resonance imaging, 1238
high-field resistive, 1095 magnetoencephalography, 1244–1245
hybrid, 1098 nuclear magnetic resonance, 1239–1243
permanent, 1085–1087 nuclear quadrupole resonance imaging,
polyhelix, 1097 1243–1244
pulsed, 1094, 1099–1102 Megagauss generators, 1094, 1102
state of the art, 1096 Melt spinning, 1397
superconducting, 1089 Melt-spun ribbons, 1396
Magnonics, 916 Mermin-Wagner theorem, 242
Magnons, see Spin waves Mesoscopic, 393
Major hysteresis loop, 1382, 1385 quantum tunneling, 276
Majorana bound states (MBS), 1171 Metal-insulator-semiconductor (MIS), 960
Majorana fermion, 635, 637, 638, 651, 1070 Metal-to-insulator transition (MIT), 959,
Maki parameter, 632 1184, 1185
Manganese Metallic glass, 1467
crystal structure, 668 Metallic magnetic materials
Fermi-level spin polarisation of magnetic alloys/intermetallic compounds, 693
elements, 678–685 alnicos, 698, 702–704
magnetic structure, 668 compounds, 705, 716–718, 720
rare earths, 671–677 d-d and d-p types, 706–709, 711, 712, 714
Manganese-zinc ferrite, 871 d-d and d-p types, 705
Manganite, 451, 880 energy conversion applications, 694
Manganosite, 876 ferromagnetic metallic elements, 693
Mangantites, 508 magnetocaloric intermetallics, 766–771
Many-body effects, 189, 191 morphology, 694
Many-body perturbation theory, 239 MSM, 705, 716–718, 720
Many-body problem, 190, 195 phase structures, 694
Many-electron crystal-field splittings, 124 Sm-Co magnets, 757–760
Martensite, 1390 Metallic magnetic thin films
Martensitic instability, 559 3d transition metals and alloys, 810
Mass transport, 1602, 1603 crystallographic and magnetic phases, 810
forced convection, 1603 GaAs (001) substrates, 811, 812
natural convection, 1603 GMR, 810
Nernst-Planck, 1603 magnetic elementary metals, 810
Matteucci effect, 557 magnetic properties, 810
Matthiessen’s rule, 437, 440, 443 MMFs, 810
Maximum energy product, 1385 technological applications, 810
Maxwell, James Clerk, 12 Metals, 1319, 1320
Maxwell-Ampère equation, 1443 Metastable non-equilibrium, 1383
Maxwell-Faraday equation, 1440 Metastable states, 1381
Maxwellian fields, 1313 Metrology, 1141
Maxwell’s equations, 12, 13, 348, 356, 361, M-ferrites, 875
464, 1245, 1376 Michael Faraday, 11
Maxwell stress, 1092, 1605 Micro-coils, 1654
Mean field approach, 241 Micro-electro-mechanical systems (MEMS),
Mean-field approximation (MFA), 61, 79, 83, 1129, 1130
84, 92, 223, 938–940 Micro-total analysis system (μ-TAS),
Mean-field theory, 34 1655–1657
Measurement, AC-susceptibility, 1134 Microfluidic chip, 1656
Mechanical forces, 1659 Micrographies, 698
Mechanical stimulation, 1659, 1660 Micromagnetic(s), 393, 1026, 1334,
Mechanotransduction, 1659 1335, 1374
Medical implants, 1404 domains, 497
1702 Index

Micromagnetism, 25, 43, 349, 1374, 1375, precise beam flux control, 1178
1383, 1386 pseudo-binary compounds, 1178
basics, 349–350 slow growth rate, 1178
Brown’s micromagnetic equation, 376–381 substrate temperature, 1167, 1168
characteristic length scales, 372–376 UHV environment, 1164, 1165, 1167
magnetic Gibbs free energy, 350–371 Molecular beam(s), 1165
magnetization dynamics, 382 fluxes, 1165, 1167, 1168
Microphones, 1391, 1421 Molecular field, 20
Microscopic, 392 Molecular quantum spintronics, 995, 996
equation of motion, 1299–1301 direct coupling, 996–999
Microscopic three-temperature model indirect coupling, 999–1001
(M3TM), 514, 516, 519, 520 quantum algorithms, 1001–1002
MicroSQUID, 990 Molecules, 1023
Microstructure, 1393, 1395, 1396, 1412 Molybdenum permalloy powders cores, 1483
images, 702 Moment
Microwave absorption, 1479 dipole, 1595
Microwave ferrites, 1478–1480 electron, 1594
Microwave materials, 105 spin-only, 1595, 1597
Microwave power tubes, 1419 Momentum
Microwave resonator, 1322 canonical, 1105
Microwave signal, 1327 kinetic, 1105
Mineral and metal separation, 1420 Monatomic spin-chain model, 151, 152
Miniaturised sensors, 1642 Monoxides, 876
Minnesotaite, 898, 899 Monte Carlo simulation, 839, 1276
Minor loops, 1403 MO recording, 478
Mixed-valence compounds, 862 Morin transition, 865
Mixed-valence manganites, 879, 881, 884 MO spectroscopy, 495–496
Mixed-valence oxides, 878 Mössbauer spectroscopy, 744, 754, 1034
Mn dimers, 87, 965 Motors, 1391
Mn3 (Cu(1−x) Gex )N, 553 Mott detector, 1215, 1351
Mn/GaAs (001), 814–816 Mott insulator, 860
Mn2 Ga, 76 Mott-Hubbard insulators, 66
MnAl, 76 Mott-Hubbard transition, 935
MnAs precipitation, 966 Mott-Ioffe-Regel limit, 438
MnBi, 56, 76 Mott’s two-current model, 443, 444
MnF2 , 84 Multicore particles, 1644
MnFeP1-x Asx alloys, 767 Multiferroic(s), 880, 915
MnO, 84 evolving terminology, 599
MnSi, 87 ferroelectric antiferromagnets, 601–604
Mobility edge, 438 ferroelectric ferrimagnets, 602, 604
μT −model, 517 ferroelectric ferromagnets, 599, 600
Modulation technique, 1109 ferroic crystal, 598
Molar susceptibility, 1597 ferroic properties, 598
Molecular beam epitaxy (MBE), 810, 907 as primary ferroics, 599
application, 1165 single-phase multiferroics, 599
chamber, 1165, 1166 spin filter, 458
development, 1165 Multigrain particles, 1400
drawback, 1178 Multilayer heterostructures, 1556
features, 1165 Multi-phase branching, 413
Heusler compounds, 1178 Multiple-pulse AOS mechanism, 525
history, 1165 Multiple scattering theory, 199, 226
in-situ monitoring of thin film growth, Multi-pole magnetization, 1388
1168, 1169 Multi-spin Hamiltonian, 987
molecular beam fluxes, 1167, 1168 Mumetal, 580, 1465
Index 1703

Muon spin rotation, 687 energy products, 761, 762, 765


Muscle cells, 1639 fabrication, 762
Musculoskeletal imaging, 1647 iron-based hard magnets, 760
Myoglobin, 1643, 1646 magnetic materials, 761, 764
Miscut angle, 1157 nanocomposite structures, 765
Nd2 Fe14 B, 761
powder metallurgy processing, 763
N processing method, 764
Nanocolumns, 967 room temperature phase diagram, 762
Nanocomposites, 1048 TEM, 766
Nanocrystalline alloys, 1470–1472 X-ray composition micrograph, 765
amorphous alloys, 695 Nd2 Fe14 B, 84, 1397–1401, 1650, 1653
anisotropy energy, 697 Nd-Fe-B, 1397, 1399
atomic moments, 697 Neckham, Alexander, 6
B-H hysteresis loop, 695 Neodymium-iron-boron (NdFeB), 1086
bulk amorphous alloys, 696 Nerve cells, 1639
conventional melt spinning technique, 696 Net shape, 1400
core losses, 695 Neural stem cells, 1660
crystallization, 695 Neutron scattering, 24, 41
elements, 695 Neutron spallation sources, 41
ETMs, 696 New-type magnetic memories
Fe-based alloys, 696 2D van der Waals magnets memory,
Fe-Si-based, 697 1572, 1573
glass-forming, 696 antiferromagnetic memory, 1570, 1571
hysteretic loss, 698 domain wall racetrack memory,
magnetocrystalline anisotropy, 697 1565–1567
magnetostrictive coefficient, 698 skrymion racetrack memory, 1567–1569
metallic glasses, 695 Ni2 MnGa, 559
microstructure, 696 Ni/GaAs (001), 824–826
MTM, 696 Nickel (Ni), 148, 156, 1647
random anisotropy model, 697 Nickel-type anisotropy, 109
soft magnetic materials, 698–700 Nickel-zinc ferrite, 871
Nanocrystalline materials, 1387 Nickelates, 880
Nanocrystalline Nd-Fe-B, 1399 NiO, 84
Nanomagnet logic, 1574–1576 Ni-rich alloys, 1463–1465
Nano/microdisc arrays, 1060 Nitrides, 895, 896
Nanoparticles, 166, 169 Nitrobenzene, 1612
Nanopores, 1058 Nitrogen-vacancy (NV) centre, 1539
Nanoscale, 160, 167 diamond sensors, 1642
Nanoscale zero valence iron (nZVI), 1620 magnetometry, 1218, 1224
Nanostructures, 165, 167 Noise
Nanowires, 1058 1/f , 1108
edge states, in 2D topological Barkhausen, 1108
insulators, 1060 burst, 1108
template-grown, 1058, 1059 definitions, 1541, 1542
National Institute of Standards and Technology Johnson, 1108
(NIST), 1143 Poisson, 1108
Natural philosophy, 7 random telegraph, 1108
Navier-Stokes equation, 1603 shot, 1108
Nd-Fe-B permanent magnets sources, 1542
alloy compositions, 760 thermal, 1108
anisotropy, 765 white, 1108
bulk magnets, 765 Non-adiabatic torque, 428
coercivity, 762 Non-aqueous syntheses, 1018, 1019
1704 Index

Non-biomedical Nucleation, 1608, 1615


core-shell NPs, 1032, 1033 centers, 1156
L10 FePt nanoparticles, 1030, 1031 controlled magnets, 1387, 1394
magnetic recording media, 1030, 1031 of reverse domains, 1398
spin canting, 1033–1035 threshold, 1018
surface effects, 1033–1035 Néel, Louis, 23
Non-centrosymmetric superconducting Néel-Brown model, 1015, 1035, 1036
monolayers, 638 Néel’s pair-interaction model, 162, 163
Non-collinear spin configuration, 219 Néel state, 81, 93
Non-destructive evaluation, 1530 Néel temperature, 24, 1276
Non-equilibrium, 1381
Non-Fermi liquid (NFL), 772
O
Non-linear optical susceptibility, 481
Occupational exposure, 1667, 1670
Non-local, 1384
Octahedral intersticies, 1410
dissipation of angular momentum, 529
Oersted, Hans-Christian, 10
Non-oriented silicon steels (NOSS), Off-axis holography, 1212–1213
1455, 1456 Ohm’s law, 436, 1441, 1442
cutting, 1456 Oleic acid, 1034
improved texture, 1460 Olivines, 897, 898
properties, 1457 One-electron Green function, 195
thin gauge steels, 1456 One-electron approximation, see Independent-
Non-resonant electromagnetic probing, electron approximation
1127–1129 One-electron model, 853
Non-uniform FFLO state, 630, 632 One-electron wave functions, 57–59
Non-uniform spin wave modes, 1315 Onsager’s relation, 436
Nonlocal resistivity measurement, 466 Open bore MRI, 1648
Nonmagnetic ligands, 112 Operation temperature, 1394, 1397,
Nonorthogonality catastrophe, 56, 58 1398, 1413
Nonreciprocal, 478, 482 Operator equivalents, 136, 137
Nontronite, 899 Optical conductivity, 486
Nonuniform magnetic fields, 1418–1421 Optical pulse-probe methods, 42
Nonuniformly magnetized bodies, 1378 Optical second-harmonic generation, 481
Nuclear adiabatic demagnetisation (NDR), Optical skin depth, 480, 489
1497, 1498 Optical spin-transfer torque, 530–532
Nuclear magnetic moment, 1310 Optical susceptibility, 481
Nuclear magnetic resonance (NMR), 41, 1026, Optical THz spin wave excitation, 532–534
1087, 1089, 1090, 1122–1127, Optical transitions, 485
1238–1241, 1298, 1646 Optically detected magnetic resonance
detection principle, 1125 (ODMR), 1328, 1329
dynamic range, 1126 Optically induced intersite spin transfer
functional magnetic resonance (OISTR), 507
imaging, 1243 Optically-induced spin transfer, 528–530
imaging and pulse sequences, Optically-induced spin transport, 510
1241–1244 Optically pumped magnetometers, 1538, 1539
instumentation and cost, 1127 Opto-magnetism, 479, 502
signal properties and precision, Orbital angular momentum, 982
1123–1125 Orbital magnetic moments, 1374
spatial variation of external magnetic Orbital magnetization density, 207
field, 1126 Orbital moment, 106, 116, 118, 131, 133, 148,
spectroscopy, 1601 149, 157, 171, 177
time variation of external magnetic Orbital momenta, 486
field, 1126 Orbital order, 856, 857, 880
tracking range, 1126 Orbital polarization, 155
Nuclear spins, 275 Order selecting aperture (OSA), 1232
Index 1705

Organic magnetoresistance (OMAR), 454 Pauli spin susceptibility, 209


Organic-metallic interfaces, 1221 Pauli susceptibility, 626
Organic superconductors, 632, 633 Pauli’s spin matrices, 148
Orgel diagram, 126, 127 Pd-series anisotropy, 148
Orientation factor, 1259 p−d Zener model, 949–950
Orthochromites, 880 Peclet number, 1604
Orthodontics, 1650 Percolation, 911
Orthoferrites, 872–873, 880 Percolation threshold, 850, 851, 901, 902, 911
Orthoferrosilite, 898, 900 Peregrinus, Petrus, 7
Orthogonal spin-transfer (OST) driven Periodic table, 851
magnetic switching, 1342–1344 Peritectic phase formation, 1401
Orthotitanates, 880 Permalloy, 404, 508, 580, 665, 1463
Orthovanadates, 880 Permalloy Fe20 Ni80 , 27
Oscillatory behavior, 1556 Permanent magnet (PM), 1374, 1387, 1413
Overlap integral, 58 Permanent magnet materials (PMM),
Oxic-anoxic transition zone, 1636 1387–1389, 1401–1405, 1409–1412
Oxide anisotropies, 145 Permanent magnets, 4, 15, 27–30, 40, 86,
Oxide heterostructures, 914–916, 1057 1085, 1649
Oxide interfaces, 915–916 (Ba, Sr)Fe12 O19 , 156
Oxide magnetism, 852–863 electro-mechanical applications, 1087
Oxide monolayers, 909–914 Fe-C (steel), 156, 157, 166
4f oxides, 888, 889 magneto-mechanical applications, 1087
5d oxides, 887 Nd2 Fe14 B, 142
5f oxides, 889–890 Nd2 Fe14 B, 106, 114, 132, 138
Oxide spin electronics, 915 physical and material properties,
Oxide thin films, 903 1085–1086
magnetic oxide monolayers, 909–914 Sm2 Fe17 N3 , 132
oxide heterostructures and interfaces, SmCo5 , 105, 106
914–916 SmCo5 , 142
substrates, caps and buffers, 903–904 static magnetic field sources, 1086–1087
thin film preparation, 904–909 Permeability, 1439
Oxide tunnel barriers, 914 complex, 1449
Oxygen effective, 1481
magnetic phase diagram, 686–687 initial, 1439
molecule-based magnets vs. single- static, 1450
molecule magnets, 685 Permendur, 665, 666
molecule-based moments, 686 Permindur, 1461
transport, 1642 Perovskite, 448
Perovskite solid solutions, 881–884
Perovskites, 878–879
P double perovskites and related materials,
Paleomagnetism, 1374 884–885
Paramagnetic effect, 626 perovskite solid solutions, 881–884
Paramagnetic metals, 582 rare earth orthoferrites and related
Paramagnetism, 12, 20, 75, 84, 1643 compounds, 879–881
Paramagnons, 75, 95 Perovskites growth, 1185
Partial density of states (PDOS), 662 Perpendicular magnetic anisotropy, 1054
Particle-induced x-ray emission (PIXE), 964 Perpendicular standing spin waves
Pascal’s constants, 1598 (PSSW), 285
Pauli exclusion principle, 351 Perpetual motion, 7
Pauli matrices, 650 Perturbation theory, 147–149
Pauli paramagnetic effect, 631, 633 Pharmacokinetics, 1645
Pauli principle, 55, 62, 68, 858 Phase coherence time, 994
Pauli spin matrices, 21 Phase diagram, 884
1706 Index

Phase separation effects, 964–967 Pressure


Phase theory, 393 magnetic, 1091
Phase transition, 727 Projector-augmented-wave method
Phase-locked epitaxy (PLE), 1168 (PAW), 197
Phase-sensitive detection (PSD), 1133, 1323 Propagating waves
Phase-shift theory, 941 in 1D magnetic structures, 302
Phosphates, 902 Pry and Bean model, 1444
Photo-elastic modulator (PEM), 491 Pseudo-binary compounds, 1178
Photoemission electron microscopy (PEEM), Pt, 86
1233–1235 Pt-series anisotropy, 148
Photon field, 504, 511 PtCo, 84
Physical vapor deposition, 1179 Ptychography, 1233, 1236
Piezoelectric effect, 596 p-type dilute ferromagnetic semiconductors,
Piezoelectric/ferroelectric dielectric, 612 928–929, 957–960
Pinning controlled magnets, 1387 Pulsed-field magnetometer, 1134–1135
Pinning of domains walls, 1387, 1390 Pulsed inductive microwave magnetometry
Planar Hall effect, 442 (PIMM), 1327
Planck constant, 350 Pulsed laser deposition (PLD), 811, 904, 908
Planck’s constant, 21 ablated particles, 1179
PLD targets, 1181 ablation process, 1182
Pnictides, 895–896 advantage, 1179, 1180
Point contact Andreev reflection (PCAR), 678, applications, 1184, 1185
679, 682, 684 configuration, 1180
Point-charge model, 123, 139 Excimer lasers, 1181, 1182
Poisson equation, 952 film properties, 1179
Poisson ratio, 568, 578 growth of perovskites, 1185
Polar catastrophe, 910, 916 growth of VO2, 1184, 1185
Polar geometry, 485 high energy density, 1179
Polarization analysis, 1034 optics, 1183
Polarization modulation, 490, 491 plume, 1179
Polarization modulation scheme, 500 substrate heating, 1184
Polarized neutron reflectometry, 1261 superconductivity, 1179
Polarized neutron techniques, 1260–1261 targets, 1181
Polarizing beam splitter (PBS), 490, 491 thickness monitoring, 1183, 1184
Poles, magnetic, 9 Pulsed magnetic fields, 1388
Polycrystalline cubic materials, 564–565 Pulsed magnets, 1094, 1099–1102
Polycrystalline hexagonal materials, 566 Pulse-length dependence, 518
Polyhelix magnets, 1097 Pump-probe scheme, 499
Polymer bonded, 1400, 1404 p−wave superconductor, 638
Polymer bonding, 1400 Py/GaAs (001), 826–828, 830
Position sensing, 1545–1548 Pyrite, 892, 893
Post-assembly magnetizing, 1388 Pyrochlore, 885–886, 1061
Post-sinter annealing (PSA), 1398 Pyrochlore-ordered dysprosium titanate, 97
Post-sintering heat treatment, 1394, 1398 Pyrochorite, 869
Post-transition metals (PTM), 696 Pyromagnetic effect (PME), 1511–1513
Potential exchange, 937 Pyrope, 871
Powder magnetoresistance (PMR), 454 Pyrophyllite, 899
Powder metallurgy, 1394, 1395 Pyroxenes, 900
Powder metallurgy processing, 763 Pyrrhotite, 894
Power conversion
PME, 1511–1513
Power spectral density (PSD), 1542 Q
Preisach model, 1381, 1382 Quadratic coefficients, 640
Pressless process, 1398 Quadrupole field, 1418, 1419
Index 1707

Quality factor, 409 Racetrack memory, 526


Quantum algorithms, 1001–1002 Radial texture, 1388
Quantum anomalous Hall effect (QAHE), 1172 Radiation, 1024
Quantum coherence, in molecular Radical-pair based magnetoreception,
magnets, 994 1637–1639
Rabi oscillations, 994, 995 Radical-pair mechanism, 1637, 1638
resonant photon absorption, 994 Radiofrequency (RF), 1238, 1239
Quantum corral, 1069 magnetic field, 1646
Quantum critical points (QCP), 772, 774 Raising operators, see Spin-flip operators
Quantum information processing (QUIP), 275 Raman scattering, 1267
Quantum Ising Spin Networks, 263–265 Random anisotropy, 116, 169–171
Quantum magnetism atomic-scale, 170
decoherence, 265–268 intergranular exchange, 170
large-scale coherence and entanglement, magnetic hysteresis, 170
271–276 magnets, 171
quantum relaxation, in dipolar nets, net anisotropy, nanoparticles, 169
268–271 Random-phase approximation (RPA), 73
spin paths and spin phase, 262–265 Random phase approximation Green function
Quantum materials, 1563, 1564 (RPA-GF), 241
Quantum-mechanical Ising models, 164 Random telegraph noise, 1542
Quantum mechanics, 1300 Rapid quenching, 36, 1387, 1388
Quantum mirage, 1069 Rare earth, 228, 671, 889, 1375, 1394
Quantum optimisation, 271 balance, 1406, 1407
Quantum phase interference, 992–993 crystal field, 672–673
Quantum phase transitions (QPT), 772, 773 elements, 717, 1394
Quantum relaxation, in dipolar nets, 268–271 iron garnets, 872
Quantum spin Hall effect, 462 magnetic properties, 674
Quantum spin Hall systems, 1171 magnetic structures and phase
Quantum spin liquids, 93, 97 transitions, 677
Quantum theory, 29 magnetoelastic effects, 673–675
Quantum transport, 454–456 magnets, 1397–1401
Quantum tunelling, 1070 materials, 1389
Quantum tunneling of magnetization, 980, moments and exchange integrals, 675–677
988–989, 1382 nitrides, 896
LZS model, 989–991 orthoferrites, 873, 879–881
quantum phase interference, 992–993 Rare-earth (RE) anisotropy
spin parity, 991–992 rare-earth ions, 133
Quantum well (QW), 965, 1053 RE-TM anisotropies, 132
Quarter-wave plate (QWP), 490, 493–495, 502 Rare-earth crystal-field parameters, 123
Quartz clocks, 1424 Rare-earth 4f multipole moments, 139
Quartz crystal microbalance (QCM), 1184 Rare-earth (RE) ions
Quasiparticle(s), 66, 73, 78, 94 4f ions, 135
energy, 636 Eu, 133
Quenching, 68, 69, 117, 131, 137, 145, 149, Gd, 133
156, 171, 172, 174 Rare-earth magnetism, 69, 576
of the angular momentum, 1307 Rare-earth permanent magnets, 32
3d orbital moment, 131 Rare-earth transition-metal (RE-TM)
free-standing monatomic nanowire, 172 alloys, 521, 1374–1401
anisotropy contributions, 144
ferrimagnetic alloys, 495
R intermetallics, 138, 142
Rabi frequency, 994 Rashba effect, 47, 128, 429
Rabi oscillations, 994, 995, 1000, 1300 Rashba-Edelstein effect, 1563, 1570
Racah operators, 672 Raw materials, 1405
1708 Index

Rayleigh criterion, 497 RF (radio frequency) magnetron sputtering,


Rayleigh law, 1440 1190, 1191
Rayleigh loops, 1439–1440 Rhodochrosite, 902, 903
RDLM method, 246–248 Rigid spin approximation (RSA), 222,
Reactive magnetron sputtering, 1190, 1191 231–232, 240–242
Reactive oxygen species (ROS), 1021, 1666 Rigidity of magnetization, 1414
Reactive sputtering, 1190, 1191 Ritz method, 380–381
Real-time MRI, 1647 RKKY coupling, 1556, 1557
Recoil permeability, 1413 Rock magnetism, 1374
Recoil product, 1413 Room temperature
Recycling potential, 1408 adiabatic temperature change data, 1499
Red blood cells, 1643 antiferromagnetic coupling, 1498
Redox processes, 1642 Curie temperature, 1501
Reflection high energy electron diffraction Fe-Rh, 1505–1507
(RHEED), 812, 1163, 1164, 1168, gadolinium metal, 1499
1169, 1183, 1184 gas-based cooling, 1498
Regenerative medicine, 1659, 1663 Gd5 (Si,Ge)4 system, 1501–1503
Relativistic effects, 201–204 Heusler compounds, 1506, 1508
Relativistic effects in magnetism, 55, 87 isofield total entropy curves, 1499
Relaxation frequency, 1450 La(Fe,Co,Mn,Si)13 Hy system, 1504, 1505
Relaxation time, 437, 1382 (Mn,Fe)2 (P,Si) system, 1503
Remanence, 1384, 1388, 1394, 1399 magnetic cooling engines, 1509–1511
enhancement, 1388, 1403, 1404 magnetic refrigerants, 1498
Remanent magnetization, 1391 magnetothermal hysteresis, 1498
Replica symmetry breaking (RSB), 956 martensitic materials, 1506, 1508
Resistive evaporation, 906 peak-shaped response curves, 1500
Resistive solenoids, 1088 phase transition, 1499
Resistivity tensor, 436 requirements, 1498
Resolution, 1108 superparamagnets (SPMs), 1498
Resonance testing device designs, 1500
electron paramagnetic resonance Rotary swaging, 1399
(EPR), 1123 Rotating MOKE (ROTMOKE)
frequency, 1479 measurements, 822
nuclear magnetic resonance (NMR), 1087, Rotational eddy currents loss, 1448
1089, 1090, 1122–1127 Rotational hysteresis loss, 1448
Resonant coil circuit, 1531 RPA-CPA, 242
Resonant inelastic x-ray scattering (RIXS), Rudderman-Kittel-Kasuya-Yosida (RKKY)
1267–1268, 1287–1289 approximation, 675
Resonant magnetic x-ray diffraction technique, Ruddleston-Popper phases, 887
1262–1265 Ruderman-Kittel exchange, 71–72
Resonant photon absorption, 994 Ruderman-Kittel-Kasuya-Yosida (RKKY)
Resonant scattering, 934 exchange interaction, 35
Resonant tunneling, 456 Ruderman-Kittel-Kasuya-Yosida (RKKY)
Resource criticality, 1405 interaction, 772, 1052
Resource criticality assessment, 1405 Runge-Zwicknagel density functional, 78
Responsivity, 1108 Russell-Saunders coupling, 129, 130
Rest potential, 1610 Rutherford backscattering (RBS), 964
Retarded single-particle Green function, 198
Reticuloendothelial system (RES), 1021
Reynolds number, 1604
magnetic, 1606 S
R2 Fe14 B, 140 Samarium-cobalt magnets, 1086
R2 Fe17 , 140 Satellite spin, 267
R2 Fe17 N3 , 140 Saturation magnetization, 351
Index 1709

Savart, Félix, 10 SQUID, 1119


Scalar potential, 9 TMR, 1119
Scanning electron microscopy (SEM), 1140, Servomotor, 1388
1213–1216 Sesquioxides, 876–877, 889
Scanning electron microscopy with Shallow donor, 932
polarization analysis (SEMPA), Shape anisotropy, 160, 161, 1314, 1315,
1141, 1214–1216, 1351 1390, 1402
Scanning Kerr microscopy, 497 Sheet silicates, 898
Scanning probe microscopy, 1217–1219 Shen Kua, 6
MFM, 1223–1226 Sherrington-Kirkpatrick model, 956
SP-STM, 1219–1223 Sherwood number, 1604
Scanning tips, 1220 Shimming, 1649
Scanning transmission electron microscopy Shot noise, 1108, 1542
(STEM), 1031 Shubnikov-de Haas effect, 438
Scanning transmission x-ray microscopy Shutters, 1189, 1190
(STXM), 1233 SI units, 12, 49
Scanning tunneling microscopy (STM), 996, Side-jump scattering, 459
1068, 1070, 1163, 1219 Siderite, 902, 903
Scanning X-ray transmission microscopy Signal-to-noise ratio (SNR), 1541
(STXM), 1345 Silicate(s), 897–902
Scattering processes, 512 garnets, 901
Scattering tensor, 1264 Silicon steel
Schmidt number, 1604 enriched, 1460
Schrödinger equation, 188, 989 fully-processed, 1453, 1454, 1456
Screening, 65, 73, 83 semi-processed, 1453, 1454
s-d model, 517 texture, 1454
Second order perturbation theory, 446 Single atom manipulation and measurements
Second-harmonic generation(SHG), 481 anisotropy, 1070
Second-neighbour superexchange, 901 magnetism, 1070
Secondary ion mass spectrometry quantum corrals, 1069
(SIMS), 1168 Single-band Hubbard model, 645
Secular equation, 196 Single domain, 1060, 1062
Selective laser melting, 1401 Single electron spintronics, 1065, 1068
Selective laser sintering, 1401 Single electron transistor, 1065
Self-demagnetization, 1392 Single-ion anisotropy, 116, 139–141, 576
Self-energy, 78 Single-ion Ising model, 164
Self-interaction correction (SIC), 248 Single ion spin Hamiltonian, 982–986
Self-pumping behavior, 1029 Single magnetic excitations, 1267
Semiclassical model, 438 Single-molecule magnets (SMMs), 980–982,
Semimagnetic semiconductors, 929 984–986, 989, 990, 995, 999
Sendust, 581, 1483 Single nuclear spin, 997, 999
Sensitivity, 1107, 1529 Single-particle equations, 190, 191
Sensor Single-particle problem, 190
AMR, 1118 Single-turn coils, 1102–1103
arrays, 1642 Sintered ferrites, 1473–1474
Faraday rotator, 1127–1129 Sintering, 1397
GMI, 1119 Site preference, 855
GMR, 1117, 1119 Skew-scattering, 459
Hall, 1115, 1119 Skin effect, 421
induction and fluxgate, 1111–1115 Skyrmion(s), 47, 89, 1056, 1222, 1224,
magnetic, 1106 1229, 1236
mechanical and MEMS-based, 1129–1130 racetrack memory, 1567–1569
nuclear magnetic resonance (NMR), Slater determinants, 61, 63, 68, 72, 81
1122–1127 Slater-Pauling behavior, 828
1710 Index

Slater-Pauling curve, 211–213, 217 alloy and spin-disorder scattering, strong


Sm2 Co17 , 1395-1397 coupling, 942
Sm-Co magnets, 757–760 alloy and spin-disorder scattering, weak
Small angle magnetization rotation coupling, 941–942
(SAMR), 571 bound magnetic polarons, 942–944
Small angle neutron scattering (SANS), 1257, quantum localization and mesoscopic
1273, 1275 phenomena, colossal
measurements, 1032 magnetoresistance, 945
Small-bore systems, 1649 spin-splitting of extended states, strong
SmCo5 , 84, 1394-1395, 1650, 1653 coupling, 940
Smythite, 894 spin-splitting of extended states, weak
Snoeck’s limit, 1451 coupling, 938–940
Soft ferrites Specific absorption rate (SAR), 1024, 1664
microwave ferrites, 1478–1480 Spectral analysis, 1541
Mn-Zn ferrites, 1474–1477 Speedometers, 1420
Ni-Zn ferrites, 1477 Speromagnet, 869
Ni-Zn-Cu ferrite, 1478 Speromagnetism, 36
sintered ferrites, 1473 Spherical harmonics, 114, 122, 131,
spinel ferrites, 1472–1473 174, 175
Soft magnetic composites (SMC), 1483 Spherical model (n = ∞ model), 92
Soft magnetic materials, 580 Spin accumulation, 1065, 1066
cobalt based amorphous alloys, 1470 Spin angular momentum, 350
eddy currents, 1440–1444 Spin blockade, 456
electrical steels, 1452–1461 Spin caloritronics
excess loss, 1444–1448 antiferromagnet insulator NiO, 1516
high frequency losses, 1449–1451 Cartesian components, 1515
hysteresis loss, 1438 efficiency of conversion, 1514
iron and low carbon steels, 1452 electron bands, 1518
iron based amorphous alloys, 1468 ferromagnetic semiconductor, 1516
iron-cobalt alloys, 1461–1463 fluctuation-dissipation theorem, 1515
iron-nickel alloys, 1463–1467 magnetic-field dependence, 1515
lumped gap, core with, 1481 magnetic heterostructures, 1513
nanocrystalline alloys, 1470–1472 magneto-electric and thermal effects,
Rayleigh loops, 1439–1440 1517, 1518
rotational losses, 1448–1449 non-trivial magnetic anisotropy, 1516
soft ferrites, 1472–1480 quality factor, 1515
spread gap, cores with, 1481–1483 SOC, 1513
Steimetz model, 1438 spin Seebeck effect, 1514
Soft magnets, 105, 166 spin transport boundary conditions, 1517
Soft micro-magnets, 1654 temperature difference, 1515
Soft steels temperature scales, 1515
electrical steel, 664 thermal gradient, 1513
Goss texture, 664, 665 thermal spin fluctuations, 1514
permalloy, 665 transverse electric field, 1514
permendur, 665, 666 Spin canting, 1033–1035
Solenoid Spin-charge separation, 94, 95
resistive and superconducting, 1088–1089 Spin current(s), 459–461, 479, 527–534
Solid solution, 850, 851, 870, 871, 881, 886, inverse spin Hall effect, 462
894, 897, 898, 901, 917 quantum spin Hall effect, 462
Solvay conference, 23 spin Hall effect, 461–462
Sommerfeld’s fine structure constant, 55 Spin current logic
sp-d(f ) exchange interactions all spin logic, 1578, 1581
sp-d(f ) exchange interactions and AND/OR gate, 1581
spin-orbit coupling, 945–946 COPY logic operation, 1581
Index 1711

graphene spin logic, 1581, 1582 Spin mixing, 514


magnetoelectric spin-orbit logic, 1583, Spin Nernst effect (SNE), 1517
1584 Spin-orbit coupling, 67, 105, 127, 128, 133,
nonlocal geometry, 1581 188, 202, 208, 215, 221, 223, 225,
Spin density functional theory, 189–195 446, 486, 487, 514, 574, 635, 829,
Spin density waves (SDWs), 669, 816, 1273 859, 888, 890, 1055, 1057, 1070,
commensurate SDWs, 670 1513, 1597
Spin-dependent scattering, 1557 of d electrons, 129
Spin-dependent two-photon of f electrons, 130
photoemission, 513 Russell-Saunders coupling, 129, 130
Spin detector, 1215 Spin-orbit interaction, 23, 47, 680, 983, 1307,
Spin-diffusion, 447 1308, 1335, 1374, 1375, 1394
Spin diffusion length, 44 Spin-orbit matrix elements, 149–150
Spin disorder scattering, 442 Spin-orbit quenching, 127
Spin dynamics Spin-orbit torque, 47, 526
with neutrons, 1282–1287 Spin-orbit torque magnetic random-access
with RIXS, 1287–1289 memories (SOT-MRAMs),
Spin echo, 1242 1562–1564
Spin electronic(s) Spin parity, 991–992
anisotropic magnetoresistance, 44 Spin polarization, 452, 453
conventional electronics, 43 Spin polarization, of 3d elements, 680
diffusion length, 44 field emission, 682
electron spin transport, 43 later photoemission experiments, 680, 681
exchange bias, 44 spin-polarised photoemission, 680
GMR, 44 Spin polarization, of 4f elements
MTJ, 45 field emission, 684
Rashba effect, 47 heavy rare earths, 684–685
sensors, 1536, 1537 PCAR, 682
spin Hall effect, 47 SXPS, ARUPS and SARPES, 683
spin-orbit interaction, 47 TM, 683
spin-orbit torque, 47 Spin-polarized angle resolved photoemission
spin transfer torque, 47 spectroscopy (SP-ARPES), 1237
spin valves, 44 Spin-polarized current, 426
thin film heterostructures, 44 Spin-polarised photoemission, 680
TMR, 45 Spin-polarized scanning tunneling microscopy
Spin filter, 457, 914 (SP-STM), 1217, 1219–1223
Spin-flip operators, 89 Spin pumping, 916
Spin-flip scattering, 443, 514 Spin quantum number, 352
Spin glasses, 35, 86, 165, 169, 170 Spin reorientation transition (SRT), 168,
Spin-glass systems, 952–957 837, 1475
Spin Hall effect (SHE), 47, 283, 428, Spin-resolved density of states, 488
429, 461–462, 916, 1327, Spin-resolved two-photon photoemission
1343, 1513 experiments, 505
Spin-Hall nano-oscillator (SHNO), 324, 325, Spin SEM, 1214
328, 329 Spin–spin interactions
Spin Hamiltonian, 63 carrier-mediated spin-spin coupling,
giant spin, 988 949–952
multi-spin, 987 dipole-dipole interactions, 946
single ion, 982–986 direct, 946
Spin ice, 97, 886, 1061–1063 superexchange, 947–948
artificial, 1062 Spin spiral calculations, 233–236
Spin lifetime, 1065, 1066 Spin-state transition, 559
Spin liquid, 880, 886, 888 Spin structure, 165
Spin magnetization, 191 Spin susceptibility, 244
1712 Index

Spin temperature, 513 Spinel (MgAl2 O4 ) structure, magnetite, 24


Spin-torque nano-oscillator (STNO), 283 Spinodal decomposition (SD), 698, 1390
and emitted spin waves, 321 Spinodal nanodecomposition, 936, 965–966
Spin torque oscillator (STO), 432, 1035 Spintronics, 276
spin transfer induced excitation of SPLEED detector, 1215, 1216
spin-waves, 1344–1346 Splitting, 854, 855
spin transfer vortex oscillators, 1346–1347 Spontaneous magnetization, 1374, 1388, 1395
Spin-transfer torque (STT), 47, 283, 426, 428, Spontaneous magnetostrictive
430, 431, 479, 530, 532, 1059, 1070, deformation, 367
1337, 1344, 1345, 1583 Spontaneous volume magnetostriction, 553
anti-damping magnetic switching, Spontaneously broken symmetry, 1380
1339–1342 Spring magnets, 1403
in phenomenological form, 1337–1339 Sputtering, 907
Spin-transfer torque magnetic random-access Sr2 Fe12 O19 , 1653
memory (STT-MRAM), 1560, 1562 S-shaped magnetization, 1026
Spin transfer vortex oscillators, 1346–1347 Stable structure phases, 814
Spin triangle, 96 Stainless steels, 666–667
Spin-valve, 447, 1050 Standard
Spin wave(s), 23, 84, 89–95, 1283–1285, 1479 primary, 1141
backscattering process, 290 working, 1142
BLS experimental technique, 288 Standard Reference Materials (SRM® ), 1143
complimentary approach, 300 State-filling effects, 507
concept of, 282 t2g states, 577
in 0D, 320–339 Static applications, 1412
in 1D waveguides, 291, 311–315 Static electric field, 436
in 3D and 2D, 284 Static magnetic field sources, 1648, 1649
dispersion curves, 236, 295, 298 Steel magnets, 1388, 1390
eigenmodes, 231 Steimetz model, 1438
energy, 238 Steinmetz exponent, 1438
excitations, 233, 237 Stephens operators, 673
frequency of, 310 Stepping motor, 1424, 1425
history of, 283 Steven’s coefficients, 983
in laterally confined systems, 291 Stevens operators, 983
lateral quantization of, 291 Stoner criterion, 914
micro-focus BLS, 300 Stoner enhancement factor, 209
propagating waves, 307 Stoner exchange integral, 209, 210
pseudo-color logarithmic maps, 322 Stoner model, 60, 208–211, 520
quantitative analytical description, 298 Stoner-Wohlfarth model, 36
resonance, 1317, 1318 Strain dependence, 578
scattering process, 289 Strain-magnetization coupling, 613
spatial and temporal evolution, 283 Stratonovich-Hubbard functional integral
spectrum, 223 method, 242
spin-torque transfer effect and, 315–320 Stray field, 14, 1376–1378, 1381, 1384, 1415
stiffness constant, 223 energy, 401
theory of, 284 mapping, 1139–1141
typology of, 288 Stress annealing, 1471
use of, 339 Stress-induced magnetic anisotropy,
waveguide, 309 567–569
wavevector, 310 Strongly exchange-enhanced Pauli
wells and edge modes, 296 paramagnets, 75
with backscattering geometry, 293 Strontium ferrite, 1392
Spindle motor, 1423 Sublattice effect, 86
Spinel(s), 869–871 Submicron range, 1399
ferrites, 30, 1472–1473 Substitution potential, 1407
Index 1713

Substrate(s), 903–905 Susceptibility, 20, 23, 74, 189, 904, 1595


heating, 1184 Curie law, 1596, 1597
temperature, 1167, 1168 diamagnetic, 1595, 1598
(Sub)THz standing spin waves, 532 dimensionless, 1596, 1598
Sucksmith-Thompson method, 115 molar, 1598
Sum-rules, 496 of liquids, 1598
Summation rules, 486 one molar, 1598
Sunspot cycle, 17 paramagnetic, 1595
Superconducting coils, 1648, 1649, 1653 Pascal’s constants, 1598
Superconducting gap, 627, 628 tensor, 237
Superconducting magnets, 42, 43, 1089
s−wave superconductor, 635
Superconducting permanent magnet,
Switchable magnets, 1419
1375, 1420
Switching efficiency, 1342
Superconducting quantum interference
Switching schemes, 1348–1349
device (SQUID), 271, 825, 1014,
Symmetric Néel wall, 416
1119–1122, 1218, 1244, 1532, 1533 Symmetry arguments, 1376
DC, 1120–1122 Synchronous motor, 1422
RF, 1121, 1122 Synchrotron, 1231
sensors, 609, 1641, 1642 sources, 42
SQUID-based MCG, 1641 Synthetic antiferromagnet, 45
Superconductivity, 1057, 1062 micro-discs, 1660
antiferromagnetic order, 640–648 Synthetic-ferrimagnetic multilayer, 526
ferromagnetic superconductors, 648–651 Spin Seebeck effect (SSE), 1513, 1514
Ising superconductors, 638–639 Step-Flow growth, 1158, 1160
paramagnetic limit and non-uniform FFLO Stranski-Krastanov growth, 1158
superconducting state, 628–634
type-I behavior, 519
type-II behavior, 519 T
Zeeman field and spin-orbit interaction, Tanabe-Sugano diagram, 126, 127
635–638 Tedrow-Messervey (TM), 678, 683, 684
Superconductor, 559, 1057, 1062 Temperature
Superdiffusive, 529 dependence of anisotropy, 141–144
spin currents, 1352 Temperature coefficient of resistivity
Superexchange, 30, 69–71, 947 (TCR), 436
antiferromagnetic, 948 3-Temperature model (3TM), 505
ferromagnetic, 948 Template-grown nanowires, 1058, 1059
interaction, 861 Tenorite, 876
IV-VI DMSs with rare earth metals, 948 Tensor notation, 586
Superhyperfine (SHF) interaction, 1310 Tephroite, 902
Superlattice, 1052 Terfenol-D, 674
Superparamagnetic fluctuations, 1036 Tetragonal L10 phases FePt and CoPt, 1404
Superparamagnetism, 36, 1060, 1067 Tetragonal magnets, 141
Surface aggregation, 967 Tetragonality, 1390
Surface anisotropies, 167, 168 Tetramethyltetraselenafulvalene
Surface coating, 1023 (TMTSF), 632
Surface diffusion, 1156 Texture, 1399, 1412
Surface effects, 1033–1035 cubic, 1460
Surface kinetics, 1155–1159 fiber, 1454
Surface magneto-optic Kerr effect Goss, 1457
(SMOKE), 497 Theorems of Kohn and Hohenberg, 190
Surface orientation, 410 Theory of Curie temperature, 950–952
Surface preparation, 1158, 1159 Thermal alloys, 1466
Surface spins, 1034 Thermal cracking, 1156
Surface state, 1056–1058, 1069 Thermal equilibrium, 1380, 1381
Surface unit cells, 1162 state, 1326
1714 Index

Thermal evaporation, 906–907 THz magnonics, 534


Thermalization time, 513 THz pulses, 510
Thermal magnetic noise spectroscopy, 1328 Time lag in magnetization, 1381
Thermal noise, 1542 Time-dependent density functional theory, 517
Thermal single-pulse switching, 522 Time-domain techniques, 1326, 1327
Thermoablation, 1664 Time-resolved magneto-optic Kerr effect
Thermodynamics of magnets, 1379–1384 (TR-MOKE), 1328
Thermo-magnetic treatment, 703 Time-resolved MCD, 502
Thermoremanent magnetization, 38 Time-resolved MOKE (TR-MOKE), 499–502,
Thick film technology, 1398 505, 507, 508, 510, 518, 530
Thickness monitoring, 1183, 1184 Time-resolved MSHG, 502
Thiele equation, 1347 Time-resolved photoemission, 513
Thin film(s), 664, 1049 Time-resolved XMCD, 522
bilayers and multilayers, with engineered Tissue engineering, 1659, 1662, 1663
properties, 1049, 1056 Tistarite, 876, 877
2D electron gases, 1056, 1058 Titanohematite, 871
Thin film growth, 40 Titanomagnetite, 870
atomic flux/surface temperature, 1155 Toggle switching, 504
buffer layer, 1159–1164 Tolerance factor, 879
choice and preparation of substrates, Topological crystalline insulators (TCI), 1172
1159–1164 Topological insulators (TI), 1057, 1060,
CMP, 1160 1171, 1172
concepts, 1155–1164 Topologically induced emergent
epitaxial growth, 1159 frustation, 1063
freestanding substrates, 1160 Topological superconducting state, 636
growth modes, 1155 Topology, 34
heteroepitaxial templates, Torque, 10, 226, 1652
1160, 1161 magnetic, 10
heteroepitaxy, 1160 magnetometer, 1136–1139
kinetic processes, 1155 magnetometry, 116
lattice mismatch, 1161, 1162 moment, 837
lattice-matched epitaxial growth Total angular momentum, 1595
strategies, 1163 Total conductivity tensor, 440
non-centrosymmetric crystal Total electronic energy and magnetic
structure, 1164 configuration
RHEED, 1163, 1164 exchange coupling parameters, 220–224
Step-Flow growth, 1160 magneto-crystalline anisotropy, 224–230
STM, 1163 total electronic energy and magnetic ground
surface preparation, 1159 state, 218
surface unit cells, 1162 total electronic energy and magnetic ground
vapor-solid phase transition, 1155–1159 state, 220
Thin film preparation, 904–906 Traction force microscopy, 1659
atomic layer deposition, 908–909 Transcranial magnetic stimulation (TMS),
chemical vapour deposition, 908 1667, 1668
molecular beam epitaxy, 907 Transfection, 1661
pulsed laser deposition, 908 Transfer matrix, 482, 484
sputtering, 907 Transferrin, 1643, 1645
thermal evaporation, 906–907 Transformer cores, 1459
Thin tube morphology, 1511 Transient ferromagnetic-like state, 523
ThMn12 −type materials, 1402, 1411 Transition metal(s) (TMs), 189, 195, 197,
Thouless-Kosterlitz transition, 92 203, 211, 216, 219, 225, 244,
Three-temperature-method, 1169 577–578, 947
Three-temperature model (3TM), 513–515 ions, 125, 983–984
THz emission, 508 rare-earth alloys, 508
Index 1715

Transition metal dichalcogenides (TMDC), Ultra-high vacuum (UHV), 812, 1165,


1175–1177 1167, 1216
Transition-metal anisotropies environment, 1164
CF and band structure, 150, 151 Ultra-soft magnetic materials, 580
perturbation theory, 147, 148 Ultraviolet photoemission spectroscopy
spin-orbit matrix elements, 149, 150 (UPS), 1168
Transmission electron microscopy (TEM), 42, Ulvöspinel, 870
766, 1019, 1208 Uniaxial, 1374
aberration correction, 1213 anisotropy, 365
electron holography, 1211–1213 anisotropy constants, 140
Lorentz microscopy, 1208–1211 stress, 568
Transmission x-ray microscopy (TXM), 1232 Uniformly magnetized, 1376, 1378
Transverse configuration, 485 Uniform magnetic field, 1377, 1378,
Transverse relaxation, 1647 1415–1417
Transverse resistivity, 442 Unpolarized neutron diffraction, 1272
Traveling-wave tube amplifier (TWTA), 1326 Unrestricted Hartree-Fock approximation,
Triarylmethyl radicals, 1599 62, 79
Triplet state, 1599 US Food and Drug Administration
Troilite, 894 (FDA), 1644
Tunneling dynamics, 270, 273
Tunneling magnetoresistance (TMR), 39, 45, V
46, 451–454, 456, 722, 1065–1068, Vaesite, 893
1117, 1118, 1536, 1557, 1559 Valence electrons, 1374
Tunneling magnetoresistive sensor, 1555 Van-der-Pauw-method, 465
Tunneling spins, 263 van der Waals layered structure, 1171
Two-band model, 440 Van der Waals (VdW) magnetism, 92
Two-contact configuration, 465 Vapor pressure, 1167
Two-current model, 443 adsorption, 1156
Two-dimensional electron gases (2DEG), 916, aggregation, 1156
1056, 1058, 1165 buffer layer, 1159
Two-dimensional magnetic materials, columnar growth, 1158
1169–1171 decomposition, 1156
Two-dimensional magnetism, 92 desorption, 1156
Two-dimensional phase transitions, 167 equilibrium, 1155–1159
2D topological insulators, 1060 Frank-van-der-Merve growth, 1157
2D van der Waals magnets memory, Gibbs free energy, 1155
1572, 1573 growth modes, 1155–1159
Two-ion anisotropy, 162, 164 growth morphology, 1156
Two-phase branching, 412 homoepitaxy, 1158
Two-step demagnetization, 519 island growth, 1158
Two-temperature model, 513 layer-by-layer growth, 1156, 1157
miscut angle, 1157
nucleation centers, 1156
U Step-Flow growth, 1158
Ultrafast demagnetization, 503 Stranski-Krastanov growth, 1158
Ultrafast laser-induced magnetization surface diffusion, 1156
dynamics, 502 surface kinetics, 1155–1159
all-optical switching of magnetization, surface preparation, 1158
521–527 temperature, 1156
laser-pulse-excited spin currents, 527–534 vapor pressures, 1156
ultrafast laser-induced loss of magnetic Volmer-Weber growth, 1158
order, 505–521 Vector network analyzers (VNA), 1324
Ultrafast magnetization dynamics, 497–502 Vector potential, 1105
Ultrafast spin transfer torques, 1351–1353 Verwey transition, 866
1716 Index

Very weak itinerant ferromagnets X-ray emission spectroscopy (XES), 966


(VWIFs), 74 X-ray holography, 1237
Vibrating reed magnetometer, 1140 X-ray imaging, 1227
Vibrating sample magnetometer (VSM), 42, CXI, 1235–1237
827, 1133–1134 PEEM, 1233–1235
Vienna Ab-initio Simulation Package SP-ARPES, 1237
(VASP), 154 STXM, 1233
Villari effect, 553, 556 TXM, 1232
Virgin curve, 1382 XMCD, 1227–1231
Virtual-crystal approximation (VCA), 938–940 X-ray magnetic circular dichroism (XMCD),
Viscosity coefficient, 1386 116, 213, 226, 495, 496, 502,
Vivianite, 902, 903 508, 523, 529, 1227–1231,
VO2 growth, 1184, 1185 1263, 1350
Voice coil actuator, 1421 sum rules, 213
Voice coil linear motors, 1422 X-ray magnetic linear dichroism
Voigt notation, 586 (XMLD), 1263
Volmer-Weber growth, 1158 X-ray microscopy, 1348, 1349
Volume magnetostriction, 550, 552, 553, X-ray photoemission spectroscopy
567, 576 (XPS), 1168
Vortex, 1060 X-ray techniques, 495
Vortex-state micro-discs, 1660 XY model, 92
Voxel, 1245

Y
W YCo5 , 148, 157
Wagner-Mermin theorem, 94 Young’s modulus, 554, 578
Walker breakdown, 424 Yttrium iron garnet (YIG), 30, 871,
Wannier functions, 58 916, 1480
Waste-water treatment, 1619
Weak localization, 439
Weak skin effect approximation, 1440 Z
Weiss, Pierre, 20, 1436, 1461 Zeeman effect, 634
Weiss field, 240 Zeeman energy, 26, 350, 357, 360, 400, 407,
Weiss model, 514 421, 452, 635
Weiss’ molecular field, see Mean-field Zeeman interaction, 55, 64, 68, 83
approximation (MFA) Zeeman splitting, 631
Whole-body MRI, 1648, 1649 Zeeman torque, 423, 424
Wide-field Kerr microscopy, 497 Zero bias anomaly (ZBA), 1068
Wiedemann effect, 556, 557, 571 Zero field resistivity, 450
Wiggler, 1418, 1419 Zero field splitting (ZFS), 983, 984,
Wind turbines, 1423 988, 1308
Working point, 1412 Zero-magnetostriction alloys, 580–582
World Health Organisation (WHO), 1666 Zhang-Rice polaron, 932
Wüstite, 867–868, 876 Zinc ferrite, 870
Zn bonding, 1404
X Zone plate, 1232
X-ray absorption (XAS), 1263 Z-phase, 1395
X-ray diffraction (XRD), 966 ZrZn2 , 56, 74

You might also like